18.11.2012 Views

full paper - RosDok - Universität Rostock

full paper - RosDok - Universität Rostock

full paper - RosDok - Universität Rostock

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Leibniz–Institut für Katalyse e.V. an der <strong>Universität</strong> <strong>Rostock</strong><br />

New Aspects in Homogeneous Hydrogenation –<br />

Development of Practical Rhodium Phosphine Catalysts<br />

and Environmentally Benign Iron Catalysts<br />

Dissertation<br />

zur Erlangung des akademischen Grades eines<br />

Doktor rerum naturalium (Dr. rer. nat.)<br />

vorgelegt der<br />

Mathematisch–Naturwissenschaftlichen Fakultät der <strong>Universität</strong> <strong>Rostock</strong><br />

von<br />

Stephan Enthaler<br />

geb. am 27.02.1980<br />

<strong>Rostock</strong>, Mai 2007


Die vorliegende Arbeit entstand in der Zeit von September 2004 bis April 2007 am Leibniz–<br />

Institut für Katalyse e.V. an der <strong>Universität</strong> <strong>Rostock</strong>.<br />

Vollständiger Abdruck der von der Mathematisch–Naturwissenschaftlichen Fakultät der<br />

<strong>Universität</strong> <strong>Rostock</strong> zur Erlangung des akademischen Grades eines Doktors der<br />

Naturwissenschaft genehmigten Dissertation.<br />

Gutachter der Dissertation: Prof. Dr. Matthias Beller, Leibniz–Institut für<br />

Katalyse e.V. an der <strong>Universität</strong> <strong>Rostock</strong><br />

(Erstgutachten)<br />

Prof. Dr. Serafino Gladiali, Dipartimento di<br />

Chimica, Universita` di Sassari, Sassari, Italien<br />

(Zweitgutachten)<br />

Prof. Dr. Uwe Rosenthal, Leibniz–Institut für<br />

Katalyse e.V. an der <strong>Universität</strong> <strong>Rostock</strong><br />

(Drittgutachten)<br />

Termin des Rigorosums: 05. November 2007<br />

Prüfungsvorsitzender: Prof. Dr. Eckhard Vogel, <strong>Universität</strong> <strong>Rostock</strong><br />

Prüfer Hauptfach (Organische Chemie): Prof. Dr. Matthias Beller, Leibniz–Institut für<br />

Katalyse e. V. an der <strong>Universität</strong> <strong>Rostock</strong><br />

Prüferin Nebenfach (Toxikologie): PD Dr. Birgit Tiefenbach, <strong>Universität</strong> <strong>Rostock</strong><br />

Wissenschaftliches Kolloquium: 15. Januar 2008


Ändere die Welt; sie braucht es.<br />

Bertolt Brecht (1898-1956)


Ich danke herzlich meinem hochgeschätzten Lehrer<br />

Prof. Dr. Matthias Beller<br />

für die Aufnahme in seine Arbeitsgruppe, die ausgezeichneten Arbeitsbedingungen, seinen<br />

unerschöpflichen Ideenreichtum und sein großes Interesse am Gelingen dieser Arbeit, sowie<br />

die gewährte wissenschaftliche Freiheit und sein Vertrauen.


Weiterhin gilt mein besonderer Dank den Mitgliedern der explorativen Projektgruppe<br />

„Synthese von neuen chiralen Liganden“ die einen maßgeblichen Anteil am Gelingen dieser<br />

Arbeit hatte: Meiner Themenleiterin Frau Dr. Kathrin Junge für die ausgezeichnete und<br />

freundschaftliche Zusammenarbeit und die gewährten Freiheiten während der letzten Jahre,<br />

Dipl.–Chem. Giulia Erre für ihre Freundschaft, Zusammenarbeit, den unzähligen Korrekturen<br />

und der kritischen Durchsicht dieser Arbeit (molte grazie!), ferner Dr. Bernhard Hagemann<br />

für seine Freundschaft und die vielen „wissenschaftlichen“ Diskussionen.<br />

Ferner gilt mein Dank Monika Heyken, Caroline Voss und Christine Mewes für die<br />

Durchführung ungezählter Experimente.<br />

Allen Freunden und Kollegen danke ich für ihre Unterstützung, Diskussionen, Chemikalien<br />

und die hervorragende Arbeitsatmosphäre: Dr. Sandra Hübner, Dipl. Chem. Kristin Schröder,<br />

Dipl.–Chem. Björn „Hartmut“ Loges, Dipl.–Chem. Hanns Martin Kaiser, Dipl.–Chem.<br />

Bianca Bitterlich, Dr. Kristin Mertins, Dipl.–Chem. Jan Schumacher, Dr. Jette Kischel,<br />

Dipl.–Chem. Cathleen Buch, Dr. Andrea „Lingelchen“ Christiansen, , Dr. Dirk Strübing,<br />

Dipl.–Chem. Anne Grotevendt, Dipl.–Chem. Anne Brennführer, Dipl.–Chem. Nicolle<br />

Schwarz, Dipl.–Chem. Karolin Alex, Karin Buchholz, Sandra Giertz, Dipl.–Chem. Benjamin<br />

Schäffner, Dipl.–Chem. Christian Torborg, Dipl.–Chem. Stefan Schulz, Dr.. Stefan Klaus,<br />

Dipl.–Chem. Thomas „Bob“ Schmidt, Dr. Marko Hapke, Dr. Ralf Jackstell, Dr. Henrik<br />

Junge, Dr. Helfried Neumann, Dr. Alexander Zapf, Dr. Hajun Jiao, Dr. Irina Jovel, Dr.<br />

Thomas Schareina, Dr. Holger Klein, Dr. Alexey Sergeev, Dr. Nicolas Clement, Chem. Ing.<br />

Christa Fuhrmann, Dr. Jens Holz, Dr. Annegret Tillack und vielen mehr.<br />

Herrn Dr. Man–Kin Tse danke ich für unzählige Diskussionen und den unerschöpflichen<br />

Vorrat an neuen Liganden.<br />

Prof. Dr. Serafino Gladiali für die vielen Diskussionen und die Zusammenarbeit.<br />

Dr. Torsten Dwars für die unzähligen Diskussionen und die unkomplizierte<br />

Chemikalienbeschaffung.<br />

Dem Analytik Team des Leibniz–Instituts für Katalyse: Dr. Dirk Michalik, Dr. Christine<br />

Fischer, Dr. Anke Spanneberg, Dr. Wolgang Baumann, Susanne Schareina, Karin Buchholz,


Kathleen Mevius, Astrid Lehmann, Kathrin Reincke, Christine Domke und im Besonderen<br />

Susann Buchholz für die Messung unzähliger GC–Proben.<br />

Der Degussa Homogeneous Catalysis (DHC) für die freundliche Bereitstellung<br />

verschiedenster Liganden.<br />

Mein ganz besonderer Dank gebührt meinen Eltern für ihre Unterstützung.


Table of Contents<br />

1. Introduction<br />

Page<br />

1<br />

2. Phosphorous containing ligands in the field of asymmetric<br />

hydrogenation<br />

3<br />

2.1. Monodentate phosphine ligands in rhodium–catalyzed asymmetric<br />

hydrogenation reactions of C–C double bonds<br />

4<br />

2.2. Monodentate ligands based on 4,5–dihydro–3H–dinaphtho[2,1–c;1´,2´–<br />

e]phosphepine motif in asymmetric hydrogenations of C–C double<br />

bonds<br />

17<br />

2.2.1. Monodentate ligands based on 4,5–dihydro–3H-dinaphtho[2,1–c;1´,2´–<br />

e]phosphepine scaffold in asymmetric synthesis<br />

29<br />

2.3. Concluding remarks 30<br />

2.4. References 31<br />

3. Reduction of unsaturated compounds with homogeneous iron<br />

catalysts<br />

39<br />

3.1. Introduction 40<br />

3.2. Hydrogenation of carbonyl compounds 41<br />

3.3. Hydrogenation of Carbon–Carbon Double Bounds 47<br />

3.4. Hydrogenation of imines and imino–analogues compounds 55<br />

3.5. Hydrosilylation 56<br />

3.6. Concluding Remarks 63<br />

3.7. References 63<br />

4. Objectives of the work 65<br />

4.1. Application of monodentate phosphine ligands in asymmetric<br />

hydrogenation<br />

65<br />

4.1.1. Asymmetric hydrogenation of N–acyl enamides 65<br />

4.1.2. Asymmetric hydrogenation of β–dehydroamino acid derivates 65<br />

4.1.3. Asymmetric hydrogenation of enol carbamates 66<br />

4.2. Transfer hydrogenation 66<br />

4.2.1. Homogeneous iron catalyst in reduction chemistry 66<br />

4.2.2. Transfer hydrogenation of ketones with ruthenium catalysts 67


4.2.2.1. Transfer hydrogenation with ruthenium carbene catalysts 67<br />

4.2.2.2. Asymmetric transfer hydrogenation with ruthenium catalysts 67<br />

4.3. References 68<br />

5. Application of monodentate phosphine ligands in asymmetric<br />

hydrogenation<br />

5.1. Enantioselective rhodium–catalyzed hydrogenation of enamides in the<br />

presence of chiral monodentate phosphines<br />

5.2. Development of practical rhodium phosphine catalysts for the<br />

hydrogenation of β–dehydroamino acid derivates<br />

5.3. Enantioselective rhodium–catalyzed hydrogenation of enol carbamates<br />

in the presence of monodentate phosphines<br />

87<br />

6. Transfer hydrogenation 100<br />

6.1. Homogeneous iron catalyst in reduction chemistry 100<br />

6.1.1. An environmentally benign process for the hydrogenation of ketones<br />

with homogeneous iron catalysts<br />

101<br />

6.1.2. Biomimetic transfer hydrogenation of ketones with iron porphyrin 108<br />

catalysts<br />

6.1.3. Biomimetic transfer hydrogenation of 2–alkoxy– and 2–aryloxyketones<br />

with iron porphyrin catalysts<br />

118<br />

6.2. Transfer hydrogenation of prochiral ketones with ruthenium<br />

catalysts<br />

125<br />

6.2.1. Efficient transfer hydrogenation of ketones in the presence of ruthenium<br />

N–heterocyclic carbene catalysts<br />

131<br />

6.2.2. New ruthenium catalysts for asymmetric transfer hydrogenation of<br />

prochiral ketones<br />

140<br />

7. Summary 143<br />

7.1. Application of monodentate phosphines in asymmetric hydrogenations 143<br />

7.2. Transfer hydrogenation 144<br />

69<br />

69<br />

76


Abbreviations<br />

Ac Acetyl<br />

Acac Acetylacetonate<br />

Anisyl Methoxylphenyl group<br />

Ar Aryl<br />

BICHEP 2,2´–bis(diphenylphosphino)–6,6´–dimethoxy–1,1´–biphenyl<br />

BINAP 2,2´–Bis(diphenylphosphino)–1,1´–binaphthyl<br />

Binaphane 1,2–Bis[4,5–dihydro–3H–binaphtho[1,2–c:2´,1´–e]phosphepino]benzene<br />

f-Binaphane 1,1´–Bis{4,5–dihydro–3H-dinaphtho[1,2–c: 2´,1´–<br />

e]phosphepino}ferrocene<br />

Binapine 4,4´–Di–tert–butyl–4,4´,5,5´–tetrahydro–3,3´–bis-3H–dinaphtho[2,1–<br />

c:1´,2´–e]phosphepine<br />

BINEPINE 4,5–Dihydro–3H–dinaphtho[2,1–c;1´,2´–e]phosphepine<br />

BINOL 2,2´–Binaphthol<br />

Bipy-tb Bis–tert–butyl–bipyridine<br />

Bn Benzyl<br />

BMPP Benzyl methyl phenyl phosphine<br />

Bopa-ip Bis(2–((S)–4–iso–propyl–4,5–dihydrooxazol–2–yl)phenyl)amine<br />

Bopa-tb Bis(2–((S)–4–tert–butyl–4,5–dihydrooxazol–2–yl)phenyl)amine<br />

BPE 1,2-Bis(2,5–diethylphospholano)ethane<br />

BPPM Butoxycarbonyl–4–diphenylphosphino–2–diphenylphosphinomethyl–<br />

pyrrolidine<br />

t–Bu tert–Butyl<br />

i–Bu iso–Butyl<br />

Bz Benzoyl<br />

CAMP Cyclohexyl o–anisylmethylphosphine<br />

Cat. Catalyst<br />

ChiraPhos Bis(diphenylphosphino)butane<br />

cod 1,5–Cyclooctadiene<br />

coe Cyclooctene<br />

Conv. Conversion<br />

Cp Cyclopentadienyl<br />

Cp´ Substituted cyclopentadienyl<br />

Cy Cyclohexyl<br />

DABCO 1,4–Diazabicyclo[2.2.2]octane<br />

DeguPhos Bis(diphenyl-phosphino)–1–benzyl–pyrrolidine<br />

DIOP O-Isopropyliden–2,3–dihydroxy–1,4–bis(diphenylphosphino)butane<br />

L–DOPA 3,4–Dihydroxyphenylalanine<br />

DIPAMP 1,2–Bis[(2–methoxyphenyl)(phenyl)phosphino]ethane<br />

DMSO Dimethylsulfoxide<br />

DPPE 1,2-Bis(diphenylphosphino)ethane<br />

DuPhos Bis(2,5–dimethylphospholano)benzene<br />

ee Enantiomeric excess<br />

e.g. Exempli gratia<br />

equiv. Equivalent<br />

Et Ethyl<br />

et al. et alii, et aliae or et alia<br />

gem Geminal<br />

h Hour<br />

JosiPhos 1–[2–(Diphenylphosphino)ferrocenyl]ethyldicyclohexylphosphine


L Ligand<br />

MandyPhos 2,2´-Bis[(N,N-dimethylamino)(phenyl)methyl]–1,1´–<br />

bisdicyclohexylphosphino)ferrocene<br />

Me Methyl<br />

Ment Menthyl<br />

MeOH Methanol<br />

MonoPhos 3,5–Dioxa–4–phosphacyclohepta[2,1–a;3,4–a´]dinapthalen–4–<br />

yl)dimethylamine<br />

MPPP Methyl phenyl n–propyl phosphine<br />

nbd Norbornadiene<br />

NMDPP Neomenthyldiphenylphosphine<br />

NorPhos 2,3–Bis(diphenylphosphino)–bicyclo[2.2.1]hept–5–ene<br />

OAc Acetate<br />

OMe Methoxy<br />

o Ortho<br />

PAMP Phenyl o–anisylmethylphosphine<br />

Ph Phenyl<br />

i–Pr i–Propyl<br />

n–Pr n–Propyl<br />

2-PrOH 2–Propanol<br />

Pybox 2,6–Bisoxazolin–2–yl pyridine<br />

R Organic rest<br />

r.t. Room temperature<br />

SDS Sodium dodecylsulfonate<br />

SiPhos 10,11,12,13–Tetrahydrodiindeno[7,1–de:1´,7´–fg][1,3,2]dioxaphosphocin–<br />

5–dimethylaniline<br />

Terpy 2,2´:6´,2´´–Terpyridine<br />

THF Tetrahydrofuran<br />

tmeda N,N,N´,N´–Tetramethylethylendiamine<br />

TMS Trimethylsilyl<br />

TOF Turnover frequency<br />

TON Turnover number<br />

X Halide


1. Introduction 1<br />

1. Introduction<br />

The importance of chiral compounds is demonstrated by their auspicious widespread<br />

applications as building blocks or intermediates for the synthesis of pharmaceuticals,<br />

agrochemicals, polymers, natural compounds, auxiliaries, ligands and synthons in organic<br />

syntheses. 1 To serve the growing demand of enantiomerically pure products various synthetic<br />

approaches were developed. Within the diverse consumers of chiral compounds the<br />

pharmaceutical industry provides one of the main impulses for new products, since nowadays<br />

only drugs containing single enantiomers are allowed to sign up for market. 2 This<br />

requirement was initiated as a consequence of different pharmacological studies, which<br />

affirmed different behaviour of both enantiomers, and proven by accidents with racemic<br />

drugs. 3 Hence, the development of new and improved methods for the preparation of<br />

enantiomerically pure compounds is an important target of industrial and academic research.<br />

Several synthetic approaches to optically active compounds have been realized in the past.<br />

Still one of the most practical approach is the separation of enantiomers via resolution, e.g. in<br />

the presence of chiral acids or enzymatic processes. The main drawback of this concept is the<br />

poor atom efficiency and the negative environmental impact, albeit nowadays promising<br />

efforts on dynamic kinetic resolutions have been established. 4 A more attractive access is<br />

represented by applying enantiopure materials from the “chiral pool”. Another powerful<br />

strategy is the application of stereoselective syntheses in particular reactions catalyzed by<br />

transition metal complexes, biocatalysts, and in recent years by organocatalysts. 1e-g,5 Within<br />

the different molecular transformations to chiral compounds transition metal catalyzed<br />

reactions offer an efficient and versatile strategy and presents a key technology for the<br />

advancement of “green chemistry”, specifically for waste prevention, reducing energy<br />

consumption, achieving high atom efficiency and generating advantageous economics. 6 Since<br />

the mid of the last century manifold catalytic reactions have been reported, e.g.<br />

polymerization, metathesis, cross couplings, hydroformylations, aminations or epoxidations. 1g<br />

One of the most popular and extensive studied catalytic reaction with respect to industrial<br />

applications is the asymmetric hydrogenation of unsaturated compounds containing C=C,<br />

C=O and C=N bonds since addition of molecular hydrogen provides high atom efficiency and<br />

the starting material are often readily available. 7 In this regard, for activation of the molecular<br />

hydrogen or hydrogen donors and transfer of chiral information the use of transition metal<br />

catalysts containing chiral ligands is essential. Commonly, chiral phosphorous ligands were<br />

used as source of chirality since catalysts including phosphorous systems proved to be highly


1. Introduction 2<br />

active as well as enantioselective. Furthermore diverse representatives are available on large<br />

scale. 8 The relevance of hydrogenation processes are impressively confirmed by several<br />

integrations in industrial applications (Scheme 1).<br />

HO<br />

OH<br />

NH 2<br />

L-DOPA<br />

anti Parkinson´s disease<br />

(Monsanto,<br />

VEB Isis-Chemie)<br />

via C=C hydrogenation<br />

OHC<br />

OH HN O<br />

NH<br />

COOH<br />

OH<br />

1 2<br />

COOMe<br />

3 4<br />

O<br />

Et<br />

O<br />

NH t Bu<br />

Crixivan<br />

HIV protease inhibitor<br />

(Merck, Lonza)<br />

via C=C hydrogenation<br />

(-)-Menthol<br />

building block for e.g.<br />

pharmaceuticals,<br />

flavours or cosmetics<br />

(Takasago)<br />

via isomerization and<br />

C=C hydrogenation<br />

Ph<br />

precursor for<br />

penem antibiotics<br />

(Takasago)<br />

via C=O hydrogenation<br />

and<br />

dynamic kinetic resolution<br />

Citronellol<br />

fragrance, vitamine E precursor<br />

(Takasago)<br />

via C=C hydrogenation<br />

MeO<br />

S-Metolachlor<br />

herbizide<br />

(Ciba-Geigy-Syngenta-Solvias)<br />

via C=N hydrogenation<br />

OH HCl<br />

OH<br />

N<br />

Bn<br />

HO<br />

5 6 OH<br />

7<br />

Adrenaline<br />

β-adrenergic receptor activator<br />

(Boehringer-Ingelheim)<br />

via C=C hydrogenation<br />

Scheme 1. Selection of industrial important compounds synthesized via hydrogenation protocols in<br />

the presence of chiral catalysts containing chiral phosphines. 9<br />

However, the generality of asymmetric hydrogenation is deeply limited by the substrate-<br />

structure and transferability of method since no universal catalyst is on hand, which shows<br />

extraordinary behaviour for most classes of substrates. Hence the discovery of new effective<br />

ligands is an important challenge for industrial and academic research. 7,10 Furthermore the<br />

final breakthrough in industry is hampered by catalysts containing expensive and toxic late<br />

transition metals (Ir, Rh or Ru). Therefore, substitution by ubiquitous available, inexpensive<br />

and less toxic metals is one of the challenging objectives for future research. In this regard the<br />

use of iron catalysts is especially desirable. 11


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

“The inherent generality of this method offers almost unlimited opportunities for matching<br />

substrates with catalysts in a rational manner and we are hopeful that our current effort will<br />

result in real progress towards complete stereospecificity.” W. S. Knowles and M. J. Sabacky 12<br />

Since the pioneering investigations on asymmetric hydrogenation by Knowles, Kagan, and<br />

others in the late 1960s of the last century extensive efforts were undertaken to develop this<br />

powerful method. In the early stage of investigations phosphorous ligands have been proven<br />

to be beneficial by supporting the catalyst activity and transferring the chiral information. A<br />

huge number of successful ligand concepts have been developed during the last decades.<br />

Nowadays they are most frequently used in the field of asymmetric C=C bond<br />

hydrogenations. However, even if several industrial applications on small to medium scale are<br />

established the final breakthrough is so far not reached, due to the low transferability of<br />

technique to novel demands. Therefore, tailor–made catalysts designed for well–defined tasks<br />

are important tools for achieving high activity and enantioselectivity. Based on this fact the<br />

discovery of new effective ligands is a key challenge for industrial and academic research.<br />

Innovative catalysts must fulfil several requirements for successful industrial applications. On<br />

the one hand an easy to adopt synthetic approach on large scale and straightforward structural<br />

variation of the ligand key structure must be available (ligand library). Furthermore, the<br />

catalyst must attain reasonable activity and enantioselectivity in the investigated reaction. On<br />

the other hand intellectual property should be accessible. 13<br />

In this regard, the development of new ligands for asymmetric hydrogenation has focused on<br />

the synthesis of bidentate ligands for nearly 30 years. However, at the end of the 20 th century<br />

the predominant role of bidentate ligands changed and monodentate ligands received more<br />

attention. 14,15 The advantages of monodentate phosphorous ligands are their easier synthesis<br />

and tunability compared to bidentate counterparts. The capabilities of monodentate ligands<br />

were successful demonstrated in several reactions. 16 Nowadays monodentate ligands are<br />

accepted as additional tool in asymmetric hydrogenations.<br />

This review will focus on the application of monodentate phosphine ligands in asymmetric<br />

hydrogenations of C–C double bonds.<br />

3


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

2.1. Monodentate phosphine ligands in rhodium–catalyzed asymmetric hydrogenation<br />

reactions of C–C double bonds<br />

“The considerable variation of yields with phosphine structure clearly shows the need for a<br />

match of catalyst and substrate.” W. S. Knowles, M. J. Sabacky, and B. D. Vineyard 17<br />

The beginning of homogenous hydrogenation started in the mid of the sixties of the last<br />

century and is connected with the groups of Wilkinson and Vaska. Thereby the group of<br />

Wilkinson reported the first successful application of a well–defined complex [RhCl(PPh3)3]<br />

(Wilkinson´s catalyst) 18 in homogeneous hydrogenation of simple olefins with molecular<br />

hydrogen, while shortly after Vaska obtained similar results with Ir(CO)(PPh3)3 as catalyst<br />

(Vaska´s complex). 19 Based on this seminal work a first chiral version was independently<br />

presented by the groups of Knowles and Horner 20 in 1968 right after Mislow et al. 21 and<br />

Horner et al. 22 established an approach to P–chiral ligands via resolution of diastereomeric<br />

menthyl phosphinates and subsequent reaction with Grignard reagents and phosphine oxide<br />

reduction (Scheme 2).<br />

Horner and Mislow<br />

MentO<br />

Knowles and Sabacky<br />

Horner<br />

O<br />

MentO P R 1<br />

R 2<br />

COOH<br />

resolution<br />

O<br />

P R 1<br />

0.15 mol% RhCl 3(8a) 3<br />

3.5 mol% NEt 3, H 2<br />

COOH<br />

*<br />

benzene-ethanol (1:1), 60 °C<br />

9 10<br />

R<br />

R 2<br />

1. R 3 MgX<br />

2. HSiCl 3<br />

8a: R 1 = Ph, R 2 = Me, R 3 = i-Pr<br />

8b: R 1 = Ph, R 2 = Me, R 3 = n-Pr (MPPP)<br />

8c: R 1 = Ph, R 2 = Me, R 3 = Bn (BMPP)<br />

8d: R 1 = Ph, R 2 = Me, R 3 = o-Anisyl (PAMP)<br />

1/2 [Rh(cod)Cl] 2 / (8b)<br />

H 2 (1 atm)<br />

15% ee<br />

R 3 P R 1<br />

R 2<br />

benzene, r.t.<br />

11a-b 12a-b<br />

R<br />

*<br />

12a: R = Et: 7-8% ee<br />

12b: R = OMe: 3-4% ee<br />

Scheme 2. Synthetic approach to chiral monodentate phosphines and first asymmetric hydrogenations<br />

carried out with homogeneous catalyst.<br />

4


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

In the reduction of α–phenylacrylic acid 9 a promising enantiomeric excess of 15% was<br />

attained when using rhodium–complexes with ligands of type 8. Hereafter, various attempts,<br />

specially focused on modification of the ligand structure were carried out to improve the<br />

enantioselectivity but unfortunately no decisively enhancement was gained. 23 Parallel to this<br />

work Horner et al. applied an in situ catalyst composed of [Rh(cod)Cl]2 and 4 equiv. of<br />

optical active methylphenyl–n–propylphosphine 8b in the asymmetric hydrogenation of α–<br />

ethylstyrene 11a and α–methoxystyrene 11b, only low enantioselectivities (


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

to further improve the enantioselectivity in asymmetric hydrogenations. During the years<br />

various strategies for inducing the chirality were success<strong>full</strong>y demonstrated, for instance<br />

transfer of chirality by different sources, e.g. axial chirality or P–chirality. Furthermore, the<br />

bite angle model (phosphorous–metal–phosphorous angle) and the formed ring sizes of<br />

phosphorous–metal–phosphorous system displayed to be crucial. In Scheme 4 a selection of<br />

outstanding bidentate ligand systems is presented. Enantioselectivities greater then >99% ee<br />

for a tremendous number of substrates were reported.<br />

H<br />

O<br />

O<br />

H<br />

DIOP (13)<br />

(1971)<br />

BINAP (19)<br />

(1980)<br />

PPh2 PPh2 PPh2 PPh2 R PPh2 R PPh2 BIPHEP (23)<br />

(1988)<br />

Ph 2P<br />

Et<br />

PPh 2<br />

H<br />

PPh 2<br />

NorPhos (20)<br />

(1981)<br />

Et<br />

N<br />

O O<br />

BPPM (16)<br />

(1976)<br />

Fe<br />

PPh 2<br />

PPh 2<br />

PPh 2<br />

Et<br />

Et<br />

MandyPhos (24)<br />

(1989)<br />

Scheme 4. Selection of outstanding bidentate ligands. 25,27<br />

MeO<br />

P<br />

P<br />

OMe<br />

DIPAMP (17)<br />

(1977)<br />

Ph 2P PPh 2<br />

N<br />

DeguPhos (21)<br />

(1984)<br />

P P<br />

DuPhos (25)<br />

(1990)<br />

Ph 2P<br />

PPh 2<br />

Fe<br />

PPh 2<br />

ChiraPhos (18)<br />

(1977)<br />

JosiPhos (22)<br />

(1994)<br />

P P<br />

BPE (26)<br />

(1990)<br />

At the early stage of bidentate ligands, enantioselectivities up to 90% ee were obtained in the<br />

hydrogenation of phenyl alanine derivatives with monodentate CAMP ligand 29 by Knowles<br />

et al. by increasing the optical purity 28 of P–chiral ligands and the introduction of o–anisyl<br />

17, 23, 29<br />

group, which probably affords a hemilabile second coordination via methoxy–group.<br />

6<br />

PCy 2


MeO<br />

HO<br />

27<br />

2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

O Ph<br />

NH<br />

COOH<br />

0.03 mol% [Rh(1,5-hexadiene)] 2<br />

0.06 mol% 29<br />

1 equiv. NaOH, H 2 (55 psi)<br />

2-PrOH, 50 °C<br />

MeO<br />

HO<br />

*<br />

90% ee<br />

O Ph<br />

NH<br />

COOH<br />

28<br />

7<br />

P Me<br />

OMe<br />

CAMP (29)<br />

Scheme 5. Asymmetric hydrogenation of α–dehydroamino acid derivatives to yield a L–DOPA<br />

precursor.<br />

Although this excellent enantioselectivities were gained, the displacement of monodentate<br />

ligands by bidentate systems could not be impeded and led to a resting state for approximately<br />

30 years (Scheme 6). Also the pioneers of this research area directed their efforts to bidentate<br />

ligands, e.g. Knowles and co–workers synthesized a dimeric version of the PAMP ligand<br />

(Scheme 2, 8d), so called DIPAMP 17 (Scheme 4) and achieved enantioselectivities up to<br />

96% ee. 30<br />

Monodentate<br />

Ligands<br />

Bidentate<br />

Ligands<br />

Knowles,<br />

Sabacky, Horner<br />

ligand 8<br />

Kagan,<br />

DIOP<br />

Knowles,<br />

CAMP<br />

1968 1971 1972 1977 1980 1993 1994 since 2000<br />

Knowles,<br />

DIPAMP<br />

Noyori,<br />

BINAP<br />

Scheme 6. Historical classification of developments in ligand synthesis.<br />

Burk,<br />

DuPhos<br />

Togni,<br />

JosiPhos<br />

Renaissance<br />

Albeit from time to time attempts were undertaken to refocus on monodentate systems the<br />

situation did not change significantly. For instance at the end of the 70 th the ligand concept<br />

presented in Scheme 2 was taken up again by Solodar for the asymmetric hydrogenation of<br />

challenging piperitenone 30, thereby monodentate ligands induced higher enantioselectivity<br />

than bidentate systems, even if the obtained enantiomeric excesses were comparatively low<br />

(up to 38% ee) (Scheme 7). 31 The influence of reaction conditions on enantioselectivity and<br />

chemoselectivity were studied in detail but no considerable improvements were attained.<br />

Noteworthy, the chemoselectivity is crucial for this type of substrate, because as side products<br />

pulegone (hydrogenation of the endocyclic double bond), menthones (hydrogenation of both


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

double bonds) and menthols (hydrogenation of both double bonds and the ketone<br />

functionality) were observed.<br />

O<br />

[Rh(cod)(29) 2]BF 4<br />

H 2 (120 psig)<br />

MeOH, 80 °C, 3 h<br />

* O<br />

30 31<br />

38% ee<br />

conv. 72%, selectivity: 45%<br />

Scheme 7. Synthesis of piperitone 31 via asymmetric hydrogenation.<br />

P Me<br />

OMe<br />

CAMP (29)<br />

Valentine et al. used a similar ligand concept as described by Morison and co–workers 24 for<br />

the reduction of ambitious compound 32 (Scheme 8), which resembles a precursor of α–<br />

tocopherol (vitamin E) or phylloquinone (vitamin K1). 32 As additional chiral carrier to the<br />

chiral menthyl–group a P–chirality was embedded and enantioselectivities up to 79% ee were<br />

attained with rhodium catalyst [Rh(cod)(34)2]BF4. Unfortunately, no information on the<br />

outcome of the reaction was given regarding the influence of ligand stereochemistry since<br />

matched and mismatched combinations are feasible.<br />

Me<br />

Ph P<br />

Me<br />

32<br />

N<br />

H<br />

COOH<br />

0.7 mol% [Rh(cod)(34) 2]BF4, H2 (2.75 atm)<br />

COOH<br />

34<br />

COOH<br />

NHAc<br />

NaOMe (1.1 equiv.),<br />

H 2 (2.75 atm), MeOH, r.t.<br />

79% ee (R),<br />

conv. >99%, TOF = 2.4 h -1<br />

0.1 mol% [Rh(cod)Cl] 2 0.2 mol% equiv. CAMP<br />

H 2 (2.75 atm)<br />

α-Tocopherol (35)<br />

COOH<br />

Me N<br />

36 MeOH, 23 °C, 36 h<br />

H 37<br />

Scheme 8. Asymmetric hydrogenations carried out by Valentine et al.<br />

67% ee<br />

conv. >99%<br />

Furthermore, α–dehydroamino acids precursors were hydrogenated in the presence of<br />

monodentate ligands, e.g. CAMP 29, in a comparative study with DIOP 13. The obtained<br />

33<br />

O<br />

NHAc<br />

8<br />

OH


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

results led to a marginal higher enantioselectivity for DIOP (72% ee) compared to CAMP<br />

(67% ee).<br />

In the mid of the 1980s Morita et al. developed an approach to sugar–based monodentate<br />

phosphines. 33 Initially started from readily available D–glucose a series of chemical<br />

transformations were carried out to attain ligand 38a. The catalytic potential was tested in the<br />

rhodium–catalyzed asymmetric hydrogenation of N–acetyl dehydrophenylalanine and the<br />

corresponding methyl ester (Scheme 9). Excellent enantioselectivity of 92% ee was obtained,<br />

albeit long reaction times were necessary. A competitive experiment performed with DIOP as<br />

ligand lead to somewhat lower enantioselectivity of 75% ee and demonstrated the<br />

extraordinary behaviour of monodentate ligand 38a under described conditions. The<br />

unfavourable long reaction times were overcome by increasing the initial hydrogen pressure,<br />

unfortunately accompanied by degradation of enantioselectivity. Furthermore, the influence<br />

of metal–ligand–ratio was studied. A significant effect on enantioselectivity as well as<br />

conversion was found when a ratio of 1:1 was used. The enantioselectivity decreased to 31%<br />

ee and a prolonged reaction time of one week was necessary to reach <strong>full</strong> conversion. Indeed,<br />

this result was approved by previous works by Knowles et al. concerning the necessity of 2<br />

equiv. ligands with respect to the metal, when a monodentate coordination mode is present.<br />

However, the catalytic behaviour of ligand 38b, synthesized by reduction of 38a, displayed a<br />

converse performance, since utilizing a metal–ligand–ratio of 1:1 a slight increase of<br />

enantioselectivity and a comparable conversion were observed. Interestingly, also the<br />

antipode enantiomer was detected. The authors assumed a bidentate coordination of the<br />

ligand, due to the attendance of the NH2–functionality and the construction of a six–<br />

membered ring. In addition, both systems were tested in the hydrogenation of itaconic acid<br />

derivatives; thereby ligand 38a exhibited once more an extraordinary enantioselectivity up to<br />

90% ee.<br />

H<br />

NC<br />

H<br />

O<br />

PPh 2<br />

O<br />

PPh2 NH2 O<br />

O<br />

O<br />

LiAlH 4<br />

O<br />

38a<br />

38b<br />

Ph<br />

NHAc<br />

COOR<br />

Scheme 9. Sugar–based monodentate ligands by Morita et al.<br />

2 mol% [Rh(cod) 2Cl] / 38<br />

H 2 (1 atm)<br />

benzene-methanol (1:4), r.t.<br />

Ph<br />

ligand: 38a (4 mol%)<br />

15b: R = H: 92% ee (S), conv. 84% (72 h)<br />

40: R = Me: 71% ee (S), conv. 51% (72 h)<br />

ligand: 38b (2 mol%)<br />

15b: R = H: 39% ee (R), conv. >99% (72 h)<br />

40: R = Me: 16% ee (R), conv. >99% (72 h)<br />

9<br />

* NHAc<br />

COOR<br />

14b, 39 15b, 40


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

Marinetti and co–workers focused on the synthesis of chiral phosphiranes, which are<br />

interesting ligands, because of their high s–character of the phosphorous lone pair, but<br />

unfortunately accompanied by high cyclic strain, which causes easy ring opening. 34 However,<br />

insertion of optical active phosphorous compounds into enantiopure styrene oxide yields<br />

molybdenum phosphirane complexes and subsequent transmetallation with rhodium<br />

complexes attained suitable precursors for asymmetric hydrogenation. The authors carried out<br />

a matched and mismatched study of all diasteriomeric combinations of ligand 41 in the<br />

hydrogenation of N–acetyl dehydrophenylalanine methyl ester 39 (Scheme 10). The best<br />

performance was shown by ligand 41b with an enantioselectivity of 76% ee, while all other<br />

combinations gained almost racemic mixtures. Marinetti et al. assume a better transfer of<br />

chirality, due to the cis–position of the phenyl group with respect to the metal, which causes<br />

steric interactions and in consequence a formation of a more stereospecific pocket. Changing<br />

the L–menthyl by t–butyl group (41e and 41f) confirmed the assumption, because trans–<br />

system led to higher enantioselectivity.<br />

Men<br />

P<br />

t-Bu<br />

P<br />

41a: P(S)C(S) (cis)<br />

41b: P(R)C(S) (trans)<br />

41c: P(R)C(R) (cis)<br />

41d: P(S)C(R) (trans)<br />

t-Bu<br />

P<br />

41e: cis 41f: trans<br />

Scheme 10. Phosphiranes as ligand in asymmetric hydrogenation.<br />

Ph<br />

NHAc<br />

Ph<br />

*<br />

COOMe<br />

NHAc<br />

1 mol% [Rh(cod)(41) 2]PF6 H2 (3.5 bar)<br />

methanol, 25 °C<br />

COOMe<br />

39 40<br />

41a: 9% ee (+)<br />

41b: 76% ee (-)<br />

41c: 3% ee (-)<br />

41d: 6% ee (+)<br />

41e: 26% ee (-)<br />

41f: 49% ee (-)<br />

Later the same group reported about an expansion of the ring size of the ligand. 35 The<br />

synthesis of chiral phosphetanes of type 42 (Scheme 11) was carried out according to the<br />

McBride approach, by reaction of branched olefins with chlorophosphine–AlCl3 complexes.<br />

The stereochemistry was regulated by chiral induction of L–menthyl as auxiliary. The acidic<br />

protons adjacent to the phosphorous atom enable further functionalization, for instance<br />

deprotonation with n–butyl lithium and subsequent quenching with benzyl bromide yields<br />

ligand 42. Ligand 42 was subjected to hydrogenation benchmark tests, but unfortunately the<br />

standard protocol, e.g. hydrogenation of N–acetyl dehydrophenylalanine with rhodium<br />

10


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

catalyst attained only low catalytic activity (8 days reaction time) and moderate enantiomeric<br />

excess of 40% ee. After switching to iridium as metal source, improved activity after 16 h was<br />

observed, but accompanied by lower enantioselectivity (


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

Pioneering work in the second age of monodentate ligands was presented in the late 90 th of<br />

the last century by the group of Fiaud applying a monodentate phospholane ligand (R,R)–45<br />

(Scheme 13), which resembles the structural motif of DuPhos 25. 37 Albeit the group of Burk<br />

demonstrated in one of their early articles the potential of ligand substituted with methyl<br />

groups (R,R)–44 in the hydrogenation of 39 (60% ee (R)) and dimethyl itaconate (65% ee (R))<br />

with catalyst activities up to 500 h -1 , their research interest focused on the development of<br />

new chiral bisphosphines and 44 was mainly used as intermediate or building block. 38<br />

However, Fiaud obtained an enantioselectivity of 82% ee in the rhodium–catalyzed<br />

hydrogenation of N–acetyl dehydrophenylalanine methyl ester using an in situ catalyst<br />

composed of 1 mol% [RhCl(cod)]2 and 2.1 mol% corresponding ligand.<br />

P Ph Burk et al.<br />

Ph<br />

P<br />

Ph<br />

44<br />

Ph<br />

45<br />

Fiaud et al.<br />

Ph<br />

NHAc<br />

COOMe<br />

cat. / H 2 (atm.)<br />

MeOH, 20-25 °C<br />

Ph<br />

39 40<br />

Burk: 60% ee (R), TON = 1000, TOF = 333 h -1<br />

(0.01 mol% [Rh(cod)(44) 2]SbF 4)<br />

Fiaud: 82% ee (S)<br />

(in situ: 1 mol% [Rh(cod)Cl] 2, 2.1 mol% 45)<br />

Fiaud: 93% ee (S), t 1/2= 12 min<br />

(in situ: 1 mol% [Rh(cod) 2]BF 4, 2.1 mol% 45)<br />

* NHAc<br />

COOMe<br />

Scheme 13. Asymmetric hydrogenation of N–acetyl dehydrophenylalanine methyl ester with<br />

monodentate phospholanes.<br />

Later on the same group reported efforts on improving the enantioselectivity, by switching<br />

from [RhCl(cod)]2 as metal source to [Rh(cod)2]BF4 which resulted in a significant increase of<br />

enantiomeric excess to 93% ee. 39 Furthermore, the scope and limitation of ligand 45 was<br />

displayed by substrate variation, once dimethyl itaconate was hydrogenated in 55% optical<br />

yield, while for the corresponding acid a selectivity of 73% ee was attained. On the other hand<br />

N–acetyl enamides, for instance (Z)–N–(1–phenylprop–1–enyl)acetamide, were hydrogenated<br />

with good enantioselectivities up to 73% ee. In their ongoing research they became interested<br />

in stabilizing the air–sensitive phosphine by the phospholanium tetrafluoroborate counterpart<br />

and liberate the phosphine during the formation of the catalyst precursor. A slight decrease of<br />

enantioselectivity (86% ee (R)), accompanied by higher catalyst activities, was shown when<br />

applying this precursor in the hydrogenation of N–acetyl dehydrophenylalanine methyl ester<br />

with comparable reaction conditions as previous reported by them. 40<br />

12


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

At the end of the 20 th century the predominant role of bidentate ligands changed and<br />

monodentate ligands received more attention since first attempts by different groups have<br />

proven extraordinary transfer of chirality by applying monodentate phosphorous ligands<br />

41, ,<br />

based on binaphthol in asymmetric hydrogenation. 42 43 The advantages of this type of<br />

monodentate phosphorous ligands are their easier synthesis and high tunability compared to<br />

bidentate phosphines, because designated target is accomplished in commonly 1–2 step<br />

synthesis starting from low cost chiral binaphthol. Scheme 14 illustrates some examples of<br />

monodentate phosphorous ligands for asymmetric hydrogenation based on the chiral 1,1´–<br />

binaphthol skeleton. Important contributions in this area came independently from Feringa<br />

and de Vries et al. 41 (phosphoramidites 46), Reetz et al. 42 (phosphites 47), and Pringle and<br />

Claver et al. 43 (phosphonites 48). Furthermore, excellent enantioselectivity was shown by<br />

Zhou and co–workers applying monodentate spiro phosphoramidites (SIPHOS, 49) 44 and<br />

many others. 45 The rediscovery of monodentate ligands was furthermore driven by conceptual<br />

review articles by Kagan and Börner, who assumed a better transfer of the chiral information<br />

due to the catalyst structure. Until now various systems based on the preliminary structure<br />

motif (46–48) have been proven to be highly active in diverse classes of asymmetric<br />

hydrogenations with enantioselectivities up to >99% ee and a few numbers of representatives<br />

have been commercialized.<br />

O<br />

P NR1R2 O<br />

O<br />

P OR<br />

O<br />

O<br />

P R<br />

O<br />

46 47 48 49<br />

Scheme 14. Selection of successful chiral monodentate ligands in asymmetric hydrogenations.<br />

13<br />

NR1R2 O P<br />

O<br />

However, during this time also some efforts on monodentate phosphines were reported.<br />

Marinetti and co–workers adopted the synthetic strategy of DuPhos–ligands, via formation of<br />

cyclic sulfates, for the synthesis of monodentate phosphetanes 50 (Scheme 15). 46 A huge<br />

number of potential ligands were accessible. Three ligands were chosen for testing their<br />

abilities in asymmetric hydrogenation of N–acetyl dehydrophenylalanine. In comparison to<br />

former achievements with phosphetanes an improved enantioselectivity up to 86% ee was<br />

monitored, albeit the i–propyl–derivative gave only low selectivity.


R<br />

P<br />

R<br />

50a: R = Me<br />

50b: R = i-Pr<br />

50c: R = Bn<br />

2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

Scheme 15. Phosphetane ligands established by Marinetti et al.<br />

Ph<br />

NHAc<br />

1 mol% [Rh(cod)(50) 2]PF6 H2 (3-5 bar)<br />

Ph<br />

* NHAc<br />

COOH<br />

Methanol, r.t., over night<br />

COOH<br />

14a 15a<br />

50a: 80% ee, conv. >99%<br />

50b: 10% ee, conv. >99%<br />

50c: 86% ee, conv. >99%<br />

Afterwards Helmchen et al. exposed the extraordinary catalytic behaviour of<br />

oxaphosphinanes in asymmetric hydrogenation reactions (Scheme 16). 47 The ligands were<br />

synthesized by mesylation of enantiopure diolethers and subsequent reaction with<br />

dilithiophenylphosphine. Because of oxygen–sensitivity a stabilization with BH3 was<br />

appropriated. The oxaphosphinanes boranes were deprotected with DABCO (1,4–<br />

diazabicyclo[2.2.2]octane) before pre–catalyst formation was carried out. Preliminary<br />

experiments with ligand 51b pointed out moderate enantioselectivity of 47% ee for<br />

hydrogenation of N–acetyl dehydrophenylalanine. However, the authors assumed an<br />

unfavoured sterical shielding of the rhodium therefore the phenyl group was removed and<br />

substituted by hydrogen. Indeed, significant higher enantioselectivity (86% ee) and reaction<br />

rates were attained after re–starting the catalytic investigations with secondary phosphine 52b.<br />

This fact was further approved in the hydrogenation of itaconic acid 53, because 73% ee for<br />

51b and 93% ee for 52b, respectively, were achieved. Moreover, excellent enantioselectivity<br />

up to 96% ee was induced by variation of the α–position adjacent to the phosphorous atom.<br />

R P R<br />

Ph BH3 R<br />

O<br />

O<br />

P R<br />

H BH3 DABCO<br />

1. Li<br />

2. MeOH<br />

DABCO<br />

R<br />

51a: R = Me<br />

51b: R = i-Pr<br />

51c: R = Cy<br />

R<br />

O<br />

P<br />

Ph<br />

O<br />

P<br />

H<br />

R<br />

R<br />

52a: R = Me<br />

52b: R = i-Pr<br />

52c: R = Cy<br />

HOOC<br />

Ph<br />

NHAc<br />

COOH<br />

COOH<br />

Scheme 16. Oxaphosphinanes approached by Helmchen et al.<br />

53<br />

[Rh(cod)(L) 2]BF 4 / H 2 (1.1 bar)<br />

20 °C, 24 h<br />

14b 15b<br />

Ph<br />

ligand 51b: 47% ee (S) (1 mol% cat., MeOH)<br />

ligand 52b: 86% ee (S) (1 mol% cat., MeOH)<br />

ligand 52c: 90% ee (R) (1 mol% cat., CH 2Cl 2)<br />

[Rh(cod)(L) 2]BF 4 / H 2 (1.1 bar)<br />

20 °C, 24 h<br />

HOOC<br />

ligand 51b: 73% ee (R) (1 mol% cat., MeOH)<br />

ligand 52b: 93% ee (R) (0.2 mol% cat., 2-PrOH)<br />

ligand 52c: 96%ee (S) (0.2 mol% cat., 2-PrOH)<br />

14<br />

* NHAc<br />

COOH<br />

* COOH<br />

54


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

The group of Börner reported the synthesis and catalytic application of monodentate chiral<br />

phospholanes 55 (Scheme 17). 48 Two general synthetic sequences were envisaged. On the<br />

one hand, as key step a heterogeneous arene hydrogenation of dibenzophosphole acid was<br />

carried out, but unfortunately only a single product was isolated, while in total five<br />

diastereomers could be formed. The X–ray structure analysis indicated a cis–selective<br />

hydrogenation. Secondly, an improved McCormack reaction was performed to build up the<br />

basic ligand structure, and follow up chemistry yields ligand 55. Further on epimerization was<br />

investigated theoretically as well as practically to broaden the access to other diastereomers.<br />

As predicted by calculations the stereocenter adjacent to the phosphorous atom allows<br />

epimerization in the presence of base and subsequent acidification. Transformation of<br />

received acids to the corresponding secondary phosphines yields two diastereomers, which<br />

were separated via borane adduct. Unfortunately, deprotection of the single diastereomers led<br />

to epimerization of the P–chiral phosphorous atom. However, the mixture of diastereomers<br />

was tested in the hydrogenation of dimethyl itaconate and N–acetyl dehydrophenylalanine<br />

methyl ester, thereby moderate enantioselectivities were attained.<br />

P<br />

H<br />

P<br />

H<br />

55a<br />

55b<br />

55<br />

(mixture of epimers<br />

55a and 55b)<br />

MeOOC<br />

Ph<br />

COOMe<br />

MeOOC * 1 mol% [Rh(cod) 2]BF4 / 2 mol% 55<br />

H2 (1 bar)<br />

25 °C, 4 h<br />

COOMe<br />

56<br />

45% ee (R) (CH2Cl2) 52% ee (R) (MeOH)<br />

57<br />

NHAc<br />

COOMe<br />

1 mol% [Rh(cod) 2]BF 4 / 2 mol% 55<br />

H 2 (1 bar)<br />

25 °C, 4 h<br />

39 23% ee (S) (CH2Cl2) 40<br />

61% ee (S) (MeOH)<br />

Scheme 17. Secondary phosphine based on phospholane in asymmetric hydrogenation.<br />

Ph<br />

* NHAc<br />

COOMe<br />

More recently, Börner et al. reported a second type of monodentate phospholes, based on<br />

tartaric acid, which allows easy integration of chiral informations. 49 The key step was a<br />

double hydrophosphination. The obtained ligands were subjected to asymmetric<br />

hydrogenation of α– and β–dehydroamino acid derivates. Excellent enantioselectivity of 92%<br />

ee was attained for the hydrogenation of β–dehydroamino acid derivatives.<br />

15


O O<br />

O O<br />

P<br />

P<br />

2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

58a<br />

O O<br />

O O<br />

58b<br />

R 1<br />

NHAc<br />

COOR 2<br />

2 mol% [Rh(cod) 2]BF 4 / 4 mol% 58<br />

H 2 (1 bar)<br />

r.t.<br />

ligand 58a:<br />

40: R1 = Ph, R2 = Me: 76% ee (S) (MeOH, 230 min)<br />

60: R1 = H, R2 39, 59 40, 60<br />

= Me: 81% ee (S) (CF3CH2OH, 17 min)<br />

NHAc<br />

COOR 3<br />

R 1<br />

ligand 58b:<br />

40: R 1 = Ph, R 2 = Me: 65% ee (S) (THF, 55 min)<br />

60: R 1 = H, R 2 = Me: 78% ee (S) (THF, 6 min)<br />

2 mol% [Rh(cod) 2]BF 4 / 4 mol% 58<br />

H 2 (1 bar)<br />

* NHAc<br />

COOR 2<br />

CF3CH2OH, r.t.<br />

ligand 58b:<br />

62a: R3 = Me: 82% ee (S) (200 min)<br />

62b: R3 61a-b<br />

= Bn: 92% ee (S) (180 min)<br />

62a-b<br />

Scheme 18. Monodentate phospholane ligands based on tartaric acid.<br />

16<br />

NHAc<br />

* COOR3 Breit et al. reported the synthesis of chiral phosphabarrelenes 62, which contain highly<br />

pyramidalized phosphorous atoms. 50 The ligands were accessible by Diels Alder reaction of<br />

substituted phosphabenzene and benzyne and subsequent separation of isomers.<br />

Unfortunately, testing these monodentate ligands in asymmetric hydrogenation protocols only<br />

moderate enantioselectivities were obtained in the hydrogenation of itaconic acids. However,<br />

improvement of enantioselectivity up to 90% ee was attained when the phosphabarrelene 62d<br />

was reacted with compound 48 (Scheme 14, R = Cl) to yield a bidentate ligand system.<br />

R 1<br />

P<br />

Ph<br />

R 2<br />

62a: R 1 = Ph, R 2 = 2-Napthyl<br />

62b: R 1 = Ph, R 2 = 3-MeO-C 6H 4<br />

62c: R 1 = Ph, R 2 = 2-MeO-C 6H 4<br />

62d: R 1 = Ph, R 2 = 2-HO-C 6H 4<br />

HOOC<br />

COOH<br />

1 mol% [Rh(cod) 2]BF4, 2.8 mol% 62<br />

H2 (1 bar)<br />

HOOC<br />

CH2Cl2, r.t.<br />

* COOH<br />

53 54<br />

Scheme 19. Chiral phophabarrelene in asymmetric hydrogenation.<br />

(+)-62a: 31% ee (R), conv. >99% (6 h)<br />

(-)-62a: 31% ee (S), conv. >99% (4 h)


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

2.2. Monodentate ligands based on 4,5–dihydro–3H–dinaphtho[2,1–c;1´,2´–<br />

e]phosphepine motif in asymmetric hydrogenations of C–C double bonds<br />

“We can expect that they (monophosphines) will play a role of increasing importance in many<br />

aspects of organometallic catalysis.” F. Lagasse and H. B. Kagan<br />

Ligands based on the 4,5–dihydro–3H–dinaphtho[2,1–c;1´,2´–e]phosphepine scaffold 66 are<br />

known since the early 90´s (Scheme 20) when Gladiali and co–workers reported the first<br />

successful synthesis of this ligand class. 51 The synthetic route is presented in Scheme 20. As<br />

initial step 2,2´–dimethylbinapthyl was synthesized via nickel–catalyzed Kumada coupling<br />

reaction of 1–bromo–1–methylnapthaline 63 and the corresponding Grignard reagent 64.<br />

63<br />

64<br />

Br<br />

+<br />

CH 3<br />

MgBr<br />

CH 3<br />

P Ph<br />

66a<br />

NiCl 2(PPh 3) 2<br />

N Pd<br />

Cl<br />

Cl<br />

Pd<br />

N<br />

67<br />

CH3 CH3 +<br />

1. n-BuLi, KO t Bu, tmeda<br />

2. Cl 2PR<br />

N<br />

Cl<br />

Pd<br />

P Ph<br />

N<br />

Cl<br />

Pd<br />

P Ph<br />

Scheme 20. Approach to chiral phosphepines according to Gladiali et al.<br />

65<br />

68a<br />

68b<br />

DPPE<br />

DPPE<br />

66<br />

+<br />

P R<br />

17<br />

P Ph<br />

(S)-66a<br />

P Ph<br />

(R)-66a<br />

The racemic 2,2´–dimethylbinapthyl 65 was selectively double–lithiated on the methyl groups<br />

and subsequent quenched with dichlorophosphines to yield the racemic 4,5–dihydro–3H–


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

dinaphtho[2,1–c;1´,2´–e]phosphepines 66. Resolution of the enantiomers was undertaken on<br />

racemic 66a and carried out by reacting the racemic mixture with (+)–di–μ−chloro-bis[(S)–<br />

N,N–dimethyl–α–phenylethylamine–2C,–N]-dipalladium 67 52 to form diastereomeric<br />

complexes 68a and 68b. The complexes were separated via crystallization. Finally, the<br />

enantiopure monodentate phosphine 66a (Ph–BINEPINE) was liberated by reacting the single<br />

diastereomeric complex with bidentate phosphines (e.g. DPPE).<br />

The potential of obtained enantiopure ligand (S)–66a was tested in asymmetric rhodium–<br />

catalyzed hydroformylation of styrene 69 under standard conditions (Scheme 21). Good<br />

iso/n–ratio of aldehyde (70:71) was found (95:5) and even if low enantioselectivity up to 20%<br />

ee was reached until this point it was the highest enantiomeric excess induced by monodentate<br />

ligands for this type of reaction. 51,53<br />

0.2 mol% [Rh(acac)(CO) 2]<br />

1 mol% (S)-66a<br />

O<br />

P Ph<br />

CO/H2 (50 bar)<br />

*<br />

benzene, 30 °C, 3 h<br />

(S)-66a 69 70 71<br />

+<br />

iso-selectivity: 95%<br />

20% ee<br />

Scheme 21. Enantioselective hydroformylation in the presence of catalyst containing ligand 66a.<br />

Some years later the group of Stelzer reported the synthesis of secondary phosphine based on<br />

phosphepine scaffold 66 (R = H) in good yields. 54 The described achiral secondary phosphine<br />

was used as building block or precursor for various ligand systems. However, from practical<br />

point of view the approaches invented by Gladiali et al. and Stelzer et al. causes some<br />

difficulties with respect to up–scaling and industrial demands since expensive auxiliaries were<br />

used and a low overall yield of final ligand was achieved.<br />

Different attempts were reported on the synthesis of 2,2´–dimethylbinaphthyl 65 in an<br />

enantiomeric pure form. 55 On the one hand various chiral versions were described for C–C<br />

bond formation via Kumada coupling by exchange of triphenylphosphine by chiral ligands,<br />

but so far best enantioselectivity of 95% ee was reported by Ito et al. 56 Furthermore, few<br />

Suzuki coupling reactions were published with comparable low enantioselectivities (85%<br />

ee). 57 Although some excellent enantioselectivities were demonstrated by C–C coupling<br />

methods, the up–scaling of these methods is limited by the availability of applied ligands. On<br />

the other hand a more convenient two step pathway starting from enantiomeric pure 2,2´–<br />

binaphthol (98% ee) was lately established since nowadays the available of 2,2´–binaphthol is<br />

18<br />

O


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

feasible on large scale. 58,59 The updated synthesis of binaphthophosphepines starts with<br />

diesterification of enantiomerically pure 2,2´–binaphthol 72 with trifluoromethanesulfonic<br />

acid anhydride in the presence of pyridine (Scheme 22). 60 The corresponding diester was<br />

obtained in quantitative yield and subsequent nickel–catalyzed Kumada coupling with methyl<br />

magnesium bromide leads to 2,2´–dimethylbinaphthyl 65 in 95% yield. 61 Two different<br />

synthetic strategies were established to obtain 4,5–dihydro–3H–dinaphtho[2,1–c;1´,2´–<br />

e]phosphepine ligands 66. On the one hand, double metallation of 2,2´–dimethylbinaphthyl 65<br />

with n–butyl lithium in the presence of tmeda (N,N,N´,N´–tetramethylethylenediamine)<br />

followed by quenching with commercially available dichlorophosphines gives ligands such as<br />

66a (P–phenyl) and 66b (P–t–butyl) in 60–83% yield.<br />

1. n-BuLi / tmeda<br />

2. Cl 2PNEt 2<br />

69<br />

HCl<br />

P NEt 2<br />

65<br />

OH<br />

OH<br />

72<br />

(CF 3SO 2) 2O / Py<br />

NiCl 2(dppp) / MeMgBr<br />

CH 3<br />

CH 3<br />

P Cl<br />

1. n-BuLi / tmeda<br />

2. Cl 2PR<br />

RMgX or RLi<br />

Scheme 22. Synthetic approach to 4,5–dihydro–3H–dinaphtho[2,1–c;1´,2´–e]phosphepine developed<br />

by Beller et al.<br />

74<br />

P R<br />

66<br />

19


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

Both ligands have been synthesized on >10g–scale. In the second procedure the dilithiated<br />

2,2´–dimethylbinaphthyl 65 is quenched with diethylaminodichlorophosphine to produce the<br />

phosphepine 73 62 which, upon treatment with gaseous HCl is converted into 4–chloro–4,5–<br />

dihydro–3H–dinaphtho[2,1–c;1´,2´–e]phosphepine 74 in 80% yield. This enantiomerically<br />

pure chlorophosphine is easily coupled with various Grignard or lithium reagents to render a<br />

broad selection of ligand 66. The limited number of commercially available<br />

dichlorophosphines and the large diversity of Grignard compounds make the access through<br />

4–chloro–4,5–dihydro–3H–dinaphtho[2,1–c;1´,2´–e]phosphepine 74 the route of choice to a<br />

library of ligands 66 (Scheme 23). 59<br />

OCH 3<br />

CF 3<br />

OCH 3<br />

OCH 3<br />

OCH 3<br />

OCH 3<br />

F F<br />

P R<br />

F<br />

F<br />

F F<br />

66c 66d 66e<br />

66a 66f<br />

66g 66h<br />

N<br />

D<br />

D D<br />

D<br />

D<br />

t-Bu t-Bu<br />

H 3C<br />

CH 3<br />

CH 3<br />

F<br />

H 3C CH 3<br />

N N<br />

N<br />

CH 3<br />

66k<br />

66i 66j<br />

66m<br />

66n<br />

66p 66q 66s<br />

66t<br />

66r<br />

66v 66w<br />

66x 66y<br />

66z<br />

N<br />

CH 3<br />

F<br />

H 3C<br />

F<br />

CH 3<br />

CH 3<br />

20<br />

F<br />

N(CH 3) 2<br />

Schema 23. Ligand library based on 4,5–dihydro–3H–dinaphtho[2,1–c;1´,2´–e]phosphepine motif.<br />

Noteworthy, parallel to the work of Beller et al. the group of Zhang described a similar<br />

synthetic approach, but using monodentate system as intermediate since their research mainly<br />

66<br />

66b


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

focused on bidentate ligands and preliminary unpromising results with ligand 66b in<br />

asymmetric hydrogenations were obtained (Scheme 24). 63 Extraordinary enantioselectivity<br />

was obtained for bidentate ligands containing the phosphepine moiety (Scheme 24, 75–77). 64<br />

P<br />

P<br />

Binaphane (75)<br />

CH 3<br />

CH 3<br />

65<br />

NHAc<br />

1. n-BuLi<br />

2. PCl 3<br />

1. 50-60%<br />

2. 40%<br />

COOMe<br />

P Cl<br />

1 mol% [Rh(cod)]SbF 6<br />

2 mol% 66b<br />

H 2 (60 psi)<br />

MeOH, r.t., 24 h<br />

H<br />

P t Bu<br />

Bu P t H<br />

t-BuMgBr<br />

60%<br />

59 60<br />

Binapine (76)<br />

74<br />

NHAc<br />

*<br />

COOMe<br />

P<br />

6% ee<br />

Fe<br />

P<br />

P<br />

66b<br />

f-Binaphane (77)<br />

Scheme 24. Synthesis of 4,5–dihydro–3H–dinaphtho[2,1–c;1´,2´–e]phosphepine by a protocol<br />

established by Zhang et al.<br />

Nevertheless, the group of Beller applied tool box 66 in various catalytic reactions. First set of<br />

catalytic experiments with ligand library 66 were dedicated to the rhodium–catalyzed<br />

asymmetric hydrogenation of α–amino acid precursors. 59c Here, methyl (Z)–α–<br />

acetamidocinnamate 39 and methyl α–acetamidoacrylate 59 were chosen as model systems.<br />

Representative results are summarized in Table 1 and Table 2. Surprisingly, initial experiment<br />

with different solvents pointed out a benefit for toluene even if toluene is an unusual solvent<br />

for hydrogenation purpose, due to catalyst inhibition. 65 However, best enantioselectivities up<br />

to 95% ee were achieved in toluene with high catalyst activities (TOF 1000–6000 h -1 ). In<br />

some cases a slight improvement of enantioselectivity and reactivity was found when catalytic<br />

amount of tenside sodium dodecylsulfonate (SDS) were placed in the reaction (Table 1,<br />

entries 2 and 3). Analysis of presented results indicates a crucial influence of substitution<br />

pattern at the phosphorous atom. Best enantioselectivities were obtained with aryl systems,<br />

while alkyl derivatives lead to low enantiomeric excess (Table 1, entry 1). In the case of<br />

asymmetric hydrogenation of substrate 39 a detailed study of various ligands with<br />

21


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

substituted–aryl groups connected to the phosphorous demonstrated no significant change in<br />

enantioselectivity, when electron–donating or electron–withdrawing functionalities were sited<br />

in the para–position. Analogous substitution in ortho–position decreased the<br />

enantioselectivity (Table 1). For the reduction of 59 a converse behaviour was attained since<br />

substitution on the aromatic unit improved the enantioselectivity (Table 2).<br />

Table 1. Asymmetric hydrogenation of 39 in the presence of catalysts containing ligand class 66.<br />

P R<br />

66a: R = Ph<br />

66b: R = t-Bu<br />

66c: R = 4-CH 3O-C 6H 4<br />

66e: R = 3,4-(CH 3O) 2C 6H 3<br />

66i: R = 2-CH 3O-C 6H 4<br />

66n: R = 2-F-C 6H 4<br />

66q: R = 3,5-(t-Bu) 2C 6H 3<br />

66s: R = 2-Naphthyl<br />

Ph<br />

NHAc<br />

COOMe<br />

1 mol% [Rh(cod) 2]BF 4<br />

2 mol% 66<br />

H 2 (1 bar)<br />

25 °C<br />

Ph<br />

39 40<br />

Entry Ligand Solvent t/2 [min] Conversion [%] ee [%] (R)<br />

1<br />

66b Toluene 31 > 99 20<br />

2 66a Toluene 50 > 99 90<br />

3 [a]<br />

66a Toluene + SDS 33 > 99 95<br />

4 66c Ethyl acetate 4 > 99 88<br />

5 [a]<br />

66q Toluene + SDS 2 > 99 95<br />

6 66i Toluene 59 > 99 64<br />

7 [a]<br />

66n Toluene + SDS 53 > 99 91<br />

8 [a]<br />

66s Toluene + SDS 50 > 99 94<br />

9 66e Toluene - > 99 93<br />

[a]<br />

20 mol% SDS (sodium dodecylsulfonate).<br />

Table 2. Asymmetric hydrogenation of 59 in the presence of catalysts containing ligand class 66.<br />

P R<br />

66a: R = Ph<br />

66c: R = 4-CH 3O-C 6H 4<br />

66i: R = 3-CH 3O-C 6H 4<br />

66n: R = 2-F-C 6H 4<br />

66q: R = 3,5-(t-Bu) 2C 6H 3<br />

NHAc<br />

COOMe<br />

1 mol% [Rh(cod) 2]BF 4<br />

2 mol% 66<br />

H 2 (1 bar)<br />

22<br />

NHAc<br />

COOMe<br />

NHAc<br />

COOMe<br />

25 °C<br />

59 60<br />

Entry Ligand Solvent t/2 [min] Conversion [%] ee [%] (R)<br />

1 66a Toluene 18 > 99 67<br />

2 66c Ethyl acetate 2 > 99 51<br />

3 66q Toluene 1 > 99 94<br />

4 [a]<br />

66n Toluene + SDS 20 > 99 84<br />

5 [a]<br />

66i Toluene + SDS 17 > 99 86<br />

[a]<br />

20 mol% SDS (sodium dodecylsulfonate).<br />

Next the potential of ligand library 66 was tested in the asymmetric hydrogenation of<br />

dimethyl itaconate 56 (Table 3). Here, best reaction outcome was obtained in


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

dichloromethane as solvent, even if to some extend lower enantioselectivities up to 88% ee<br />

were observed.<br />

Table 3. Asymmetric hydrogenation of dimethyl itaconate 56.<br />

P R<br />

66a: R = Ph<br />

66c: R = 4-CH 3O-C 6H 4<br />

66i: R = 3-CH 3O-C 6H 4<br />

66n: R = 2-F-C 6H 4<br />

66q: R = 3,5-(t-Bu) 2C 6H 3<br />

66s: R = 2-Naphthyl<br />

COOMe<br />

COOMe<br />

1 mol% [Rh(cod) 2]BF 4<br />

2 mol% 67<br />

H 2 (1 bar)<br />

CH 2Cl 2, 25 °C<br />

56 57<br />

Entry Ligand Conversion [%] t/2 [min] ee [%] (S)<br />

1 66a > 99 56 86<br />

2 66c > 99 - 86<br />

3 66q > 99 16 83<br />

4 66n 86 16 77<br />

5 66s 29 34 70<br />

6 66i > 99 - 88<br />

COOMe<br />

23<br />

COOMe<br />

A more convincing approach to enantiopure itaconic acid was quite recently reported by<br />

Gladiali and co–workers (Scheme 25). 66 Excellent enantioselectivities up to 97% ee were<br />

attained when subjecting itaconic acid to a rhodium–catalyzed transfer hydrogenation under<br />

mild reaction conditions in the presence of well–defined cationic [Rh(nbd)(66a)2] +<br />

complexes. As suitable source of hydrogen formic acid was chosen since carbon dioxide<br />

arises as only side product. Notably, by utilizing the corresponding dimethyl ester or the two<br />

possible monomethyl esters under same conditions a decrease of enantioselectivity was<br />

noticed, accompanied by a switch of product configuration. Furthermore, a comparative study<br />

was carried out by testing several monodentate as well as bidentate ligands. The presented<br />

results pointed out the potential of this ligand since only a BINAP complex achieved similar<br />

enantioselectivity.<br />

P Ph<br />

66a<br />

COOR 1<br />

COOR 2<br />

Scheme 25. Transfer hydrogenation of itaconic acid derivatives.<br />

1.5 mol% [Rh(nbd)(66a) 2]BF 4<br />

HCOOH/NEt 3 (5:2)<br />

DMSO, 22 °C, 4 h<br />

53: R 1 = R 2 = H: 97% ee (S), conv. >99%<br />

56: R 1 = R 2 = Me: 13% ee (R), conv. 57%<br />

COOR2 COOR<br />

*<br />

1


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

The application of ligand class 66 was furthermore demonstrated in the hydrogenation of N–<br />

acyl enamide 78 by the group of Reetz (Scheme 26). 67 The obtained enantioselectivities are<br />

comparable low, but surprisingly the tert–butyl–substituted ligand 66b (conversion: 94%;<br />

enantiomeric excess: 24% (R)) showed better selectivity and conversion than the phenyl–<br />

substituted ligand 66a (conversion: 50%; enantiomeric excess: 14% (R)), while in other<br />

benchmark tests such as the asymmetric hydrogenation of α–amino acid precursors, dimethyl<br />

itaconate and β–ketoesters always the opposite behaviour was observed. However, in<br />

combination with BINOL–derived phosphites or phosphonites a significant enhancement of<br />

enantioselectivity up to 89% ee was found. This combinatorial approach displayed<br />

impressively one advantage of monodentate phosphorous ligands, as the reaction outcome can<br />

easily be tuned by different combination of ligands. 42m,68<br />

P R 1<br />

66a: R 1 = Ph<br />

66b: R 1 = t-Bu<br />

O<br />

P R<br />

O<br />

2<br />

48a: R2 = CH3 47a: R2 = OCH3 Scheme 26. Asymmetric hydrogenation of N–acyl enamides.<br />

NHAc 0.2 mol% [Rh(cod) 2]BF4 NHAc<br />

0.2 mol% L1 / 0.2 mol% L2 H2 (1.3 bar)<br />

CH2Cl2, 30 °C, 20 h<br />

78 78<br />

L 1 = L 2 = 66a: 14% ee (R), conv. 50%<br />

L 1 = L 2 = 66b: 24% ee (R), conv. >99%<br />

L 1 = L 2 = 47a: 76% ee (R), conv. >99%<br />

L 1 = L 2 = 48a: 76% ee (R), conv. >99%<br />

combinatorial approach<br />

L 1 = 66a, L 2 = 47a: 57% ee (R), conv. >99%<br />

L 1 = 66b, L 2 = 47a: 89%ee (R), conv. >99%<br />

L 1 = 66a, L 2 = 48a: 53% ee (R), conv. >99%<br />

L 1 = 66b, L 2 = 48a: 87% ee (R), conv. >99%<br />

Following the original work of Reetz et al. the group of Beller investigated the reaction in<br />

presence of chiral monodentate 4,5–dihydro–3H–dinaphtho[2,1–c;1´,2´–e]phosphepine 66 in<br />

more detail (Table 4). 59f After optimization of different reaction parameters<br />

enantioselectivities up to 93% ee were achieved with rhodium catalyst containing 2 equiv. of<br />

ligand 66a. As shown in Table 4 the enantioselectivity is largely dependent on the nature of<br />

the substituent at the phosphorous atom. Further improvement up to 95% ee, which is up to<br />

now the highest enantiomeric excess obtained with monodentate phosphine ligands for this<br />

purpose, was attained when the substrate structure was modified by introduction of electron–<br />

donating groups at the aromatic part of the substrate, while substrates containing electron–<br />

withdrawing groups were hydrogenated with somewhat lower enantioselectivities.<br />

24


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

Table 4. Extension of asymmetric hydrogenation of N–acyl enamides.<br />

P R<br />

66a: R = Ph<br />

66b: R = t-Bu<br />

66c: R = 4-CH 3O-C 6H 4<br />

66e: R = 3,4-(CH 3O) 2-C 6H 3<br />

66h: R = 4-F-C 6H 4<br />

66i: R = 2-CH 3O-C 6H 4<br />

66m: R = 3,5-(CH 3) 2-C 6H 3<br />

66p: R = 4-F 3C-C 6H 4<br />

NHAc<br />

Entry Ligand Conversion [%]<br />

1 mol% [Rh(cod) 2]BF 4<br />

2.1 mol% 66<br />

H 2 (2.5 bar)<br />

toluene, 30 °C, 6 h<br />

25<br />

NHAc<br />

*<br />

78 79<br />

ee [%]<br />

1 66a >99 93 (R)<br />

2 66p >99 72 (R)<br />

3 66h >99 88 (R)<br />

4 66m >99 87 (R)<br />

5 66c >99 93 (R)<br />

6 66i >99 63 (R)<br />

7 66e >99 90 (R)<br />

8 66b >99 16 (S)<br />

Apart from α–amino acids β–amino acids received significant attention since they are useful<br />

building blocks for pharmaceuticals. Hence the groups of Beller and Gladiali reported on the<br />

successful application of ligand class 66 in the rhodium–catalyzed asymmetric hydrogenation<br />

of β–dehydroamino acids derivatives to give optical active β–amino acids derivatives. 69 First<br />

experiments emphasized a necessity of different reaction conditions for the E– and Z–isomer<br />

(Table 5). Good enantioselectivity (79% ee (R)) for the E–isomer was obtained by using 2–<br />

propanol at 2.5 bar hydrogen pressure, while higher pressure (50 bar) in ethanol was<br />

worthwhile for the Z–isomer (92% ee (R)). Noteworthy, a switch of product configuration<br />

was observed depending on the nature of the double bond, which has been scarcely<br />

reported. 70 Furthermore, a higher reaction rate was monitored for the Z–isomer, while<br />

converse behaviour was demonstrated for most other catalysts. 71 After separate optimization<br />

of reaction parameters for both isomers the ligand library 66 was subjected to catalytic<br />

reaction. Surprisingly, in the case of hydrogenation of E–61a good enantioselectivity was<br />

achieved with alkyl ligand 66b, while for other catalytic reactions (vide supra) poor to<br />

moderate enantioselectivities were reported (Table 5, entry 6). However, in the hydrogenation<br />

of the corresponding Z–isomer the catalyst containing ligand 66b was completely inactive.<br />

Here best enantiomeric excess was reached by ligand 66a (Table 5, entry 1). The usefulness<br />

of this concept was confirmed on the hydrogenation of several β–dehydroamino acids<br />

derivatives. An improved enantioselectivity of 94% ee was attained by replacement of methyl<br />

ester by ethyl ester. Some mechanistical attempts showed the necessity of 2 equiv. ligands per


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

metal, which is in agreement with previous findings by Knowles et al. This circumstance<br />

offers the possibility for a combinatorial ligand approach. Beller and Gladiali et al. adopted<br />

this concept and combined ligand 66a with achiral phosphorous ligands. Unfortunately, no<br />

significant positive effect was attained.<br />

Table 5. Hydrogenation of E–61a and Z–61a with different 4,5–dihydro–3H–dinaphtho[2,1–c;1´,2´–<br />

e]phosphepines 66.<br />

P R<br />

66a: R = Ph<br />

66b: R = t-Bu<br />

66c: R = 4-CH 3O-C 6H 4<br />

66d: R = 3-CH 3O-C 6H 4<br />

66e: R = 3,4-(CH 3O) 2-C 6H 3<br />

66m: R = 3,5-(CH 3) 2-C 6H 3<br />

Entry Ligand Isomer [a]<br />

O O<br />

NH O<br />

[Rh(cod) 2]BF4 + 2.1 (66)<br />

NH<br />

OMe<br />

Z-61a<br />

H 2 (50 bar), EtOH, 10 °C, 24 h<br />

O O<br />

NH<br />

[Rh(cod) 2]BF4 + 2.1 (66)<br />

NH<br />

MeO O<br />

E-61a<br />

Conv. [%]<br />

H 2 (2.5 bar), 2-PrOH, 10 °C, 24 h<br />

ee [%]<br />

Isomer [b]<br />

Conv. [%]<br />

O<br />

(S)-62a<br />

O<br />

(R)-62a<br />

ee [%]<br />

1 66a E–61a >99 79 (R) Z–61a >99 92 (S)<br />

2 66m E–61a 80 69 (R) Z–61a >99 80 (S)<br />

3 66c E–61a >99 88 (R) Z–61a >99 86 (S)<br />

4 66d E–61a 91 81 (R) Z–61a 90 40 (S)<br />

5 66e E–61a >99 90 (R) Z–61a >99 89 (S)<br />

6 66b E–61a >99 87 (R) Z–61a 3 60 (S)<br />

In addition, the group of Beller carried out an enantioselective reduction of enol carbamates,<br />

which allows an alternative approach to chiral alcohols. 72 Pioneering work in the field of<br />

asymmetric hydrogenation of enol carbamates has been reported by Feringa, de Vries,<br />

Minnaard and co–workers who have scored enantioselectivities up to 98% ee with rhodium-<br />

catalysts containing monodentate phosphoramidites (MonoPhos–family). 41l,73 Utilizing<br />

compound 80 as model substrate various reaction parameters were investigated in detail and<br />

enantioselectivities up to 96% ee were achieved with an in situ catalyst composed of<br />

[Rh(cod)2]BF4 and ligand 66a. Noteworthy, high thermal stability of the catalyst in the range<br />

of 10–90 °C was noticed with respect to enantioselectivity (94–96% ee). With these suitable<br />

conditions ligand library 66 was subjected to a series of enol carbamates. Despite all attempts<br />

only 66a displayed extraordinary selectivity.<br />

OMe<br />

26<br />

OMe


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

Table 6. Hydrogenation of compound 80 with different 4,5–dihydro–3H–dinaphtho[2,1–c;1´,2´–<br />

e]phosphepines 66.<br />

P R<br />

66a: R = Ph<br />

66b: R = t-Bu<br />

66c: R = 4-CH 3O-C 6H 4<br />

66k: R = i-Pr<br />

66i: R = 2-CH 3O-C 6H 4<br />

66m: R = 3,5-(CH 3) 2-C 6H 3<br />

66p: R = 4-F 3C-C 6H 4<br />

O<br />

N<br />

O<br />

80<br />

[Rh(cod) 2]BF 4 + 2.1 (66)<br />

H 2 (25 bar)<br />

MeOH, 25 °C, 6 h<br />

Entry Ligand Conv. [%] Yield [%] ee [%]<br />

1 66a >99 >99 96 (S)<br />

2 66p 40 40 28 (S)<br />

3 66m >99 >99 66 (S)<br />

4 66c >99 >99 72 (S)<br />

5 66i >99 >99 51 (S)<br />

6 66k 40 40 12 (S)<br />

7 66c 41 41 50 (R)<br />

N<br />

O O<br />

*<br />

While the group of Beller mainly focused on the variation of substituent at the phosphorous in<br />

4,5–dihydro–3H–dinaphtho[2,1–c;1´,2´–e]phosphepines Widhalm et al. reported efforts on<br />

the α–substitution adjacent to the phosphorous atom (Scheme 27). 74 Due to the rise of new<br />

stereocenters which are closer to the metal a better transfer of chiral information was<br />

assumed. Starting from 66a, which was synthesized according to literature procedures, the<br />

corresponding sulfide 82 was prepared. In addition, a deprotonation protocol was established<br />

which offers a huge variety of mono– as well as disubstituted ligands after reduction since<br />

simple quenching with electrophils are displayed. The substitution was highly stereoselective,<br />

because only single diastereomers were obtained. With these ligands in hand a comparative<br />

study was carried out to explore the influence of α– and α,α´–substitution. In the rhodium–<br />

catalyzed hydrogenation of 14a and 39 only with the α,α´–dimethyl ligand 86b comparable<br />

enantioselectivities to unsubstituted system were reported, while monosubstituted and other<br />

disubstituted representatives lead to moderate selectivity. Interestingly, similar<br />

enantioselectivity and activity was attained by utilization of only 1 equiv. of ligand with<br />

respect to rhodium. Noteworthy, substitution gained a switch of product configuration.<br />

27<br />

81


86a: R 1 = R 2 = TMS<br />

86b: R 1 = R 2 = Me<br />

86c: R 1 = R 2 = Et<br />

86d: R 1 = R 2 = Bn<br />

86e: R 1 = R 2 = i-Pr<br />

Ph<br />

NHAc<br />

COOR 3<br />

2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

R 1<br />

P Ph<br />

84a: R 1 = TMS<br />

84b: R 1 = Me<br />

84c: R 1 = Et<br />

84d: R 1 = Bn<br />

84e: R 1 = i-Pr<br />

1 mol% [Rh]<br />

2 mol% 66a or 86b<br />

H 2 (3 bar)<br />

r.t.<br />

Raney-Ni<br />

R 1<br />

CH3 CH3 P Ph<br />

R 2<br />

86<br />

Ph<br />

65<br />

1) n-BuLi<br />

2) Cl 2PPh<br />

3) S 8<br />

Raney-Ni<br />

* NHAc<br />

COOR 3<br />

R<br />

S<br />

P<br />

Ph<br />

1<br />

Scheme 27. Synthesis of α–substituted and α,α´–substituted phosphepine ligands.<br />

S<br />

P<br />

Ph<br />

82<br />

1) t-BuLi<br />

2) (R 1)X<br />

1) t-BuLi<br />

2) R 2 X<br />

R<br />

S<br />

P<br />

Ph<br />

1<br />

R 2<br />

S<br />

P<br />

Ph<br />

R 1<br />

83a 83b<br />

85<br />

15a: R 3 = H<br />

[Rh(cod)Cl] 2/NaOCl 4 , MeOH/CH 2Cl 2<br />

ligand 66a: 91% ee (R), conv. 99%<br />

ligand 86b: 91% ee (S), conv. 99%<br />

40: R 3 = Me<br />

[Rh(cod) 2]BF 4 , 2 mol% ligand, toluene<br />

ligand 66a: 85% ee (R), conv. 99%<br />

ligand 86b: 55% ee (S), conv. 98%<br />

28


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

2.2.1. Monodentate ligands based on 4,5–dihydro–3H–dinaphtho[2,1–c;1´,2´–<br />

e]phosphepine scaffold in asymmetric synthesis<br />

Apart from numerous attempts on asymmetric hydrogenation of C–C double bonds further<br />

asymmetric catalytic applications have been demonstrated the usefulness of monodentate<br />

phosphines based on 4,5–dihydro–3H–dinaphtho[2,1–c;1´,2´–e]phosphepine motif e.g. the<br />

ruthenium-catalyzed asymmetric hydrogenation of β–ketoesters, 75 asymmetric borylation,<br />

asymmetric Suzuki coupling reactions catalyzed by palladium complexes, the Pd–catalyzed<br />

umpoled–allylation of aldehydes 76 and the Pt–catalyzed alkoxycyclization of 1,5–enynes. 77<br />

The ligand itself without any metal has been proven as efficient organocatalyst for the<br />

enantioselective acylation of diols, [3+2] cycloadditions and [4+2] annulations (Scheme<br />

27). 78<br />

CHO<br />

NHAc<br />

OH<br />

R<br />

COOMe<br />

R1 COOR<br />

48% ee 95% ee 95% ee<br />

Hydroformylation<br />

Gladiali et al.<br />

Beller et al.<br />

Hydrogenation<br />

Beller et al.<br />

Transfer Hydrogenation<br />

Gladiali et al.<br />

Hydrogenation<br />

Beller et al.<br />

ROOC<br />

Ph<br />

ROOC<br />

H<br />

HO<br />

85% ee<br />

Alkoxycyclization<br />

Michelet, Gladiali, Genet et al.<br />

Ph<br />

Ph<br />

OH<br />

Ts<br />

N<br />

Ph<br />

Allylation<br />

Zanoni et al.<br />

COOEt<br />

COOEt<br />

99% ee<br />

[4+2] Annulation<br />

of Imines with Allenes<br />

Fu et al.<br />

CO2Et O<br />

R<br />

R 1<br />

70% ee<br />

95% ee<br />

[3+2] Cycloaddition<br />

of Allenes with Enones<br />

Fu et al.<br />

Ph OH<br />

P R<br />

Ph OAc<br />

22% ee<br />

Acylation<br />

Vedejs et al.<br />

Scheme 27. Applications of ligand class 66 in asymmetric synthesis.<br />

66<br />

COOR<br />

COOR 1<br />

97% ee<br />

Hydrogenation<br />

Beller et al.<br />

Transfer Hydrogenation<br />

Gladiali et al.<br />

NHAc<br />

95% ee<br />

Hydrogenation<br />

Beller et al.<br />

R 1<br />

NHAc<br />

COOR<br />

94% ee<br />

Hydrogenation<br />

Beller et al.<br />

O<br />

O<br />

96% ee<br />

Hydrogenation<br />

Beller et al.<br />

NR 1R 2<br />

29


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

2.3. Concluding remarks<br />

The rediscovery of monodentate phosphorous ligands displayed a milestone in asymmetric<br />

hydrogenation since high activities, enantioselectivities and ligand combinations for tuning<br />

the outcome of the reaction are feasible. In the case of chiral monodentate phosphines until<br />

now only a few number of ligands have been proven to entrance enantioselectivities above<br />

90% ee. Here ligands with 4,5–dihydro–3H–dinaphtho[2,1–c;1´,2´–e]phosphepine structure<br />

motif demonstrated excellent enantioselectivities up to 96% ee in several catalytic<br />

applications. However, the discovery of new effective phosphines will be an important<br />

challenge for future research due to there superior stability in comparison to monodentate<br />

phosphoramidites, phosphites and phosphonites.<br />

30


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

2.4. References<br />

1 a) I. C. Lennon, J. A. Ramsden, Org. Process Res. Dev. 2005, 9, 110–112; b) J. M. Hawkins,<br />

T. J. N. Watson, Angew. Chem. 2004, 116, 3286–3290; c) R. Noyori, T. Ohkuma, Angew.<br />

Chem. 2001, 113, 40–75; d) M. Miyagi, J. Takehara, S. Collet, K. Okano, Org. Process Res.<br />

Dev. 2000, 4, 346–348; e) E. N. Jacobsen, A. Pfaltz, Yamamoto, H. (Eds.), Comprehensive<br />

Asymmetric Catalysis, Springer, Berlin, 1999; f) R. Noyori, Asymmetric Catalysis in Organic<br />

Synthesis, Wiley, New York, 1994; g) M. Beller, C. Bolm (Eds.), Transition Metals for<br />

Organic Synthesis, Wiley–VCH, Weinheim, 2nd Ed. 2004.<br />

2 a) W. H. De Camp, Chirality 1989, 1, 2–6; b) M. N. Cayen, Chirality 1991, 3, 94–98; c)<br />

FDA, Chirality 1992, 4, 338–340.<br />

3 a) J. M. Brown, S. G. Davies, Nature 1989, 342, 631–636; b) A. P. de Silva, Nature 1995,<br />

374, 310–311; see also: K. Roth, Chem. Unserer Zeit 2005, 39, 212–217.<br />

4 a) U. T. Strauss, U. Felfer, K. Faber, Tetrahedron: Asymmetry 1999, 10, 107–117; b) R.<br />

Noyori, M. Tokunaga, M. Kitamura, Bull. Chem. Soc. Jpn. 1995, 68, 36–56; c) R. S. Ward,<br />

Tetrahedron: Asymmetry 1995, 6, 1475–1490; d) O. Pàmies, J.-E. Bäckvall, Chem. Rev. 2003,<br />

103, 3247–3261; e) V. Ratovelomanana-Vidal, J. P. Genêt, Can. J. Chem. 2000, 78, 846–851.<br />

5 P. I. Dalko, L. Moisan, Angew. Chem. 2004, 116, 5248–5286.<br />

6 a) P. T. Anastas, M. M. Kirchhoff, Acc. Chem. Res. 2002, 35, 686–694; b) P. T. Anastas, M.<br />

M. Kirchhoff, T. C. Williamson, Appl. Catal. A: Gen. 2001, 221, 3–13.<br />

7 a) H.-U. Blaser, B. Pugin, F. Spindler, J. Mol. Catal. 2005, 231, 1–20; b) H.-U. Blaser, Chem.<br />

Commun. 2003, 293–296; c) H.-U. Blaser in “Chiral Catalysis – Asymmetric<br />

Hydrogenations” supplement to chimica oggi 2004, 4–5.<br />

8 a) D. Geffroy, F. Handcock, W. P. Hems, A. Zanotti-Gerosa in “Chiral Catalysis –<br />

Asymmetric Hydrogenations” supplement to chimica oggi 2004, 38–41; b) M. Thommen, H.-<br />

U. Blaser in “Chiral Catalysis – Asymmetric Hydrogenations” supplement to chimica oggi<br />

2004, 6–8; c) Z. Zhang in “Chiral Catalysis – Asymmetric Hydrogenations” supplement to<br />

chimica oggi 2004, 10–17; d) M. van den Berg, D. Peña, A. J. Minnaard, B. L. Feringa, L.<br />

Lefort, J. A. F. Boogers, A. H. M. de Vries, J. G. de Vries in “Chiral Catalysis – Asymmetric<br />

Hydrogenations” supplement to chimica oggi 2004, 18–21; e) C. Le Ret, O. Briel in “Chiral<br />

Catalysis – Asymmetric Hydrogenations” supplement to chimica oggi 2004, 29–32; f) I. C.<br />

Lennon, P. H. Moran in “Chiral Catalysis – Asymmetric Hydrogenations” supplement to<br />

chimica oggi 2004, 34–37.<br />

31


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

9 a) W.S. Knowles, Angew. Chem. 2002, 114, 2097–2107; b) R. Selke, H. Pracejus, J. Mol.<br />

Catal. 1986, 37, 213–225; c) R. Selke, J. Organomet. Chem. 1989, 370, 241–248; d) S.<br />

Akutagawa, Appl. Catal. 1995, 128, 171–207; e) H.-U. Blaser, Adv. Synth. Catal. 2002, 344,<br />

17–31; f) WO 1996-EP2729, WO 2000-US22976; g) J. F. McGarrity, W. Brieden, R. Fuchs,<br />

H.-P. Mettler, B. Schmidt, O. Werbitzky in H.-U. Blaser, E. Schmidt (Eds), Large Scale<br />

Asymmetric Catalysis, Wiley-VCH, Weinheim, 2003; h) S. Akutagawa, Topics in Catal.<br />

1997, 4, 271–274; i) H.-U. Blaser, F. Spindler, M. Thommen, in J. G. de Vries, C. J. Elsevier<br />

(Eds.), Handbook of homogeneous hydrogenation, Wiley–VCH, Weinheim, 2007, 1279–<br />

1324.<br />

10 X. Zhang, Tetrahedron: Asymmetry 2004, 15, 2099–2100.<br />

11 The price for iron is relatively low (current price for 1 t of iron ~300 US-dollar) in<br />

comparison to rhodium (5975 US-dollar per troy ounce), ruthenium (870 US-dollar per troy<br />

ounce) or iridium (460 US-dollar per troy ounce), due to the widespread abundance of iron in<br />

earth´s crust and the easy accessibility. Furthermore, iron is an essential trace element (daily<br />

dose for humans 5–28 mg) and plays an important role in an extensive number of biological<br />

processes.<br />

12 W. S. Knowles, M. J. Sabacky, Chem. Commun. 1968, 1445–1446.<br />

13 H.-U. Blaser, M. Studer, Chirality 1999, 11, 459–464.<br />

14 F. Lagasse, H. B. Kagan, Chem. Pharm. Bull. 2000, 48, 315–324.<br />

15 I. V. Komarov, A. Börner, Angew. Chem. 2001, 113, 1237–1240.<br />

16 T. Jerphagnon, J.-L. Renaud, C. Bruneau, Tetrahedron: Asymmetry 2004, 15, 2101–2111.<br />

17 W. S. Knowles, M. J. Sabacky, B. D. Vineyard, J. Chem. Soc. Chem. Commun. 1972, 10–11.<br />

18 a) J. F. Young, J. A. Osbourne, F. A. Wilkinson, Chem. Commun. 1965, 131–132; b) J. A.<br />

Osbourne, F. H. Jardine, J. F. Young, F. A. Wilkinson, J. Chem. Soc. (A) 1966, 1711–1732; c)<br />

P. S. Hallman, D. Evans, J. A. Osborn, G. Wilkinson, Chem. Commun. 1967, 305–306; d) L.<br />

Horner, H. Büthe, H. Siegel, Tetrahedron Lett. 1968, 4023–4026; e) C. K. Brown, G.<br />

Wilkinson, Tetrahedron Lett. 1969, 22, 1725–1726; f) R. R. Schrock, J. A. Osborn, Chem.<br />

Commun. 1970, 567–568.<br />

19 L. Vaska, R. E. Rhodes, J. Am. Chem. Soc. 1965, 87, 4970–4971.<br />

20 L. Horner, H. Siegel, H. Büthe, Angew. Chem. 1968, 80, 1034–1035.<br />

21 a) O. Korpiun, K. Mislow, J. Am. Chem. Soc. 1967, 89, 4784–4786; b) O. Korpiun, R. A.<br />

Lewis, J. Chickos, K. Mislow, J. Am. Chem. Soc. 1968, 90, 4842–4846.<br />

22 L. Horner, H. Winkler, A. Rapp, A. Mentrup, H. Hoffman, P. Beck, Tetrahedron Lett. 1961,<br />

5, 161–166.<br />

32


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

23 W. S. Knowles, M. J. Sabacky, B. D. Vineyard, Chem. Technol. 1972, 2, 590–593.<br />

24 J. D. Morison, R. E. Burnett, A. M. Aguiar, C. J. Morrow, C. Phillips, J. Am. Chem. Soc.<br />

1971, 93, 1301–1303.<br />

25 T. P. Dang, H. B. Kagan, Chem. Commun. 1971, 481–482.<br />

26 Enantioselectivity based on (R)-alanine, after hydrolysis.<br />

27 a) K. Achiwa, J. Am. Chem. Soc. 1976, 98, 8265–8266; b) M. B. Fryzuk, B. Bosnich, J. Am.<br />

Chem. Soc. 1977, 99, 6262–6267; c) M. B. Fryzuk, B. Bosnich, J. Am. Chem. Soc. 1979, 101,<br />

3043–3049; d) A. Miyashita, A. Yasuda, H. Takaya, K. Toriumi, T. Ito, T. Souchi, R. Noyori,<br />

J. Am. Chem. Soc. 1980, 102, 7932–7934; e) H. Brunner, W. Pieronczyk, B. Schönhammer,<br />

K. Streng, I. Bernal, J. Korp, Chem. Ber. 1981, 114, 1137–1149; f) U. Nagel, Angew. Chem.<br />

Int. Ed. 1984, 23, 435–436; g) R. Schmid, M. Ceregetti, B. Heiser, P. Schönholzer, H.-J.<br />

Hansen, Helv. Chim. Acta 1988, 71, 897–929; h) M. J. Burk, J. E. Feaster, W. A. Nugent, R.<br />

L. Harlow, J. Am. Chem. Soc. 1993, 115, 10125–12138; i) T. Hayashi, A. Yamamoto, M.<br />

Hojo, Y. Ito, J. Chem. Soc. Chem. Commun. 1989, 495–496; j) A. Togni, C. Breutel, A.<br />

Schnyder, F. Spindler, H. Landert, A. Tijani, J. Am. Chem. Soc. 1994, 116, 4062–4066.<br />

28 Up to 95% optical purity was attained, compared to former 69%.<br />

29 B. Bogdanović, Angew. Chem. Int. Ed. 1973, 12, 954–964.<br />

30 W. S. Knowles, M. J. Sabacky, B. D. Vineyard, D. J. Weinkauff, J. Am. Chem. Soc. 1975, 97,<br />

2567–2568.<br />

31 J. Solodar, J. Org. Chem. 1978, 43, 1787–1789.<br />

32 D. Valentine Jr., K. K. Johnson, W. Priester, R. C. Sun, K. Toth, G. Saucy, J. Org. Chem.<br />

1980, 45, 3698–3703.<br />

33 a) S. Saito, Y. Nakamura, Y. Morita, Chem. Pharm. Bull. 1985, 33, 5284–5293; b) S. Saito, T.<br />

Kato, Y. Morita, H. Miyamae, Chem. Pharm. Bull. 1985, 33, 3092–3095.<br />

34 A. Marinetti, F. Mathey, L. Ricard, Organometallics 1993, 12, 1207–1212.<br />

35 a) A. Marinetti, L. Ricard, Tetrahedron 1993, 45, 10291–10304; b) A. Marinetti, L. Ricard,<br />

Organometallics 1994, 13, 3956–3962.<br />

36 O. Riant, O. Samuel, T. Flessner, S. Taudien, H. B. Kagan, J. Org. Chem. 1997, 62, 6733–<br />

6745.<br />

37 F. Guillen, J.-C. Fiaud, Tetrahedron Lett. 1999, 40, 2939–2942.<br />

38 M. J. Burk, J. E. Feaster, R. L. Harlow, Tetrahedron: Asymmetry 1991, 2, 569–592.<br />

39 F. Guillen, M. Rivard, M. Toffano, J.-Y. Legros, J. C. Daran, J.-C. Fiaud, Tetrahedron 2002,<br />

58, 5895–5904.<br />

40 C. Dobrota, M. Toffano, J.-C. Fiaud, Tetrahedron Lett. 2004, 45, 8153–8156.<br />

33


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

41 a) L. A. Arnold, R. Imobos, A. Manoli, A. H. M. de Vries, R. Naasz, B. Feringa, Tetrahedron<br />

2000, 56, 2865–2878; b) M. van den Berg, A. J. Minnaard, E. P. Schudde, J. van Esch, A. H.<br />

M. de Vries, J. G. de Vries, B. L. Feringa, J. Am. Chem. Soc. 2000, 122, 11539–11540; c) A.<br />

J. Minnaard, M. van den Berg, E. P. Schudde, J. van Esch, A. H. M. de Vries, J. G. de Vries,<br />

B. Feringa, Chim. Oggi 2001, 19, 12–13; d) D. Peña, A. J. Minnaard, J. G. de Vries, B. L.<br />

Feringa, J. Am. Chem. Soc. 2002, 124, 14552–14553; e) M. van den Berg, R. M. Haak, A. J.<br />

Minnaard, A. H. M. de Vries, J. G. de Vries, B. L. Feringa, Adv. Synth. Catal. 2002, 344,<br />

1003–1007; f) M. van den Berg, A. J. Minnaard, R. M. Haak, M. Leeman, E. P. Schudde, A.<br />

Meetsma, B. L. Feringa, A. H. M. de Vries, C. E. P. Maljaars, C. E. Willans, D. Hyett, J. A. F.<br />

Boogers, H. J. W. Henderickx, J. G. de Vries, Adv. Synth. Catal. 2003, 345, 308–323; g) D.<br />

Peña, A. J. Minnaard, A. H. M. de Vries, J. G. de Vries, B. L. Feringa, Org. Lett. 2003, 5,<br />

475–478; h) R. Eelkema, R. A. van Delden, B. L. Feringa, Angew. Chem. Int. Ed. 2004, 43,<br />

5013–5016; i) A. Duursma, J.-G. Boiteau, L. Lefort, J. A. F. Boogers, A. H. M. de Vries, J. G.<br />

de Vries, A. J. Minnaard, B. L. Feringa, J. Org. Chem. 2004, 69, 8045–8052; j) H.<br />

Bernsmann, M. van den Berg, R. Hoen, A. J. Minnaard, G. Mehler, M. T. Reetz, J. G. de<br />

Vries, B. L. Feringa, J. Org. Chem. 2005, 70, 943–951 ; k) L. Lefort, J. A. F. Boogers, A. H.<br />

M. de Vries, J. G. de Vries, Org. Lett. 2004, 6, 1733–1735; l) L. Panella, B. L. Feringa, J. G.<br />

de Vries, A. J. Minnaard, Org. Lett. 2005, 7, 4177–4180; m) R. Hoen, J. A. F. Boogers, H.<br />

Bernsmann, A. J. Minnaard, A. Meetsma, T. D. Tiemersma-Wegman, A. H. M. de Vries, J. G.<br />

de Vries, B. L. Feringa, Angew. Chem. Int. Ed. 2005, 44, 4209–4212; n) L. Panella, A. M.<br />

Aleixandre, G. J. Kruidhof, J. Robertus, B. L. Feringa, J. G. de Vries, A. J. Minnaard, J. Org.<br />

Chem. 2006, 71, 2026–2036; o) F. Giacomina, A. Meetsma, L. Panella, L. Lefort, A. H. M. de<br />

Vries, J. G. de Vries, Angew. Chem. 2007, 119, 1519–1522.<br />

42 a) M. T. Reetz, T. Sell, Tetrahedron Lett. 2000, 41, 6333–6336; b) M. T. Reetz, G. Mehler,<br />

Angew. Chem. Int. Ed. 2000, 39, 3889–3891; c) M. T. Reetz, G. Mehler, A. Meiswinkel, T.<br />

Sell, Tetrahedron Lett. 2002, 43, 7941–7493; d) M. T. Reetz, G. Mehler, A. Meiswinkel, T.<br />

Sell, Angew. Chem. Int. Ed. 2003, 42, 790–793; e) M. T. Reetz, G. Mehler, Tetrahedron Lett.<br />

2003, 44, 4593–4596; f) M. T. Reetz, G. Mehler, A. Meiswinkel, Tetrahedron: Asymmetry<br />

2004, 15, 2165–2167; g) M. T. Reetz, X. G. Li, Tetrahedron 2004, 60, 9709–9714; h) M. T.<br />

Reetz, J.-A. Ma, R. Goddard, Angew. Chem. Int. Ed. 2005, 44, 412–414; i) M. T. Reetz, Y.<br />

Fu, A. Meiswinkel, Angew. Chem. 2006, 118, 1440–1443; j) M. T. Reetz, X. Li, Synthesis<br />

2005, 3183–3185; k) M. T. Reetz, O. G. Bondarev, H.-J. Gais, C. Bolm, Tetrahedron Lett.<br />

2005, 46, 5643–5646; l) M. T. Reetz, A. Meiswinkel, G. Mehler, K. Angermund, M. Graf, W.<br />

Thiel, R. Mynott, D. G. Blackmond, J. Am. Chem. Soc. 2005, 127, 10305–10313; m) M. T.<br />

34


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

Reetz, X. Li, Angew Chem. Int. Ed. 2005, 44, 2959–2962; n) M. T. Reetz, M. Surowiec,<br />

Heterocycles 2006, 67, 567–574.<br />

43 a) C. Claver, E. Fernandez, A. Gillon, K. Heslop, D. J. Hyett, A. Martorell, A. G. Orpen, P. G.<br />

Pringle, Chem. Comm. 2000, 961–962; b) A. Martorell, R. Naasz, B. L. Feringa, P. G.<br />

Pringle, Tetrahedron: Asymmetry 2001, 12, 2497–2499.<br />

44 a) A.-G. Hu, Y. Fu, J.-H. Xie, H. Zhou, L.-X. Wang, Q.-L. Zhou, Angew. Chem. 2002, 114,<br />

2454–2456; b) Y. Fu, X.-X. Guo, S.-F. Zhu, A.-G. Hu, J.-H. Xie, Q.-L. Zhou, J. Org. Chem.<br />

2004, 69, 4648–4655; c) Y. Fu, G.-H. Hou, J.-H. Xie, L. Xing, L.-X. Wang, Q.-L. Zhou, J.<br />

Org. Chem. 2004, 69, 8157–8160.<br />

45 M. van den Berg, B. L. Feringa, A. J. Minnaard in J. G. de Vries, C. J. Elsevier (Eds.),<br />

Handbook of homogeneous hydrogenation, Wiley–VCH, Weinheim, 2007, 995–1027.<br />

46 a) A. Martinetti, V. Kruger, F.-X. Buzin, Tetrahedron Lett. 1997, 38, 2947–2950; b) A.<br />

Marinetti, S. Jus, F. Labrue, A. Lemarchand, J.-P. Genêt, L. Ricard, Synthesis 2001, 2095–<br />

2104.<br />

47 M. Ostermeier, J. Prieß, G. Helmchen, Angew. Chem. 2002, 114, 625–628.<br />

48 N. V. Dubrovina, H. Jiao, V. I. Tararov, A. Spannenberg, R. Kadyrov, A. Monsees, A.<br />

Christiansen, A. Börner, Eur. J. Org. Chem. 2006, 3412–3420.<br />

49 V. Bilenko, A. Spannenberg, W. Baumann, I. Komarov, A. Börner, Tetrahedron: Asymmetry<br />

2006, 17, 2082–2087.<br />

50 B. Breit, E. Fuchs, Synthesis 2006, 2121–2128.<br />

51 a) S. Gladiali, A. Dore, D. Fabbri, O. De Lucchi, M. Manassero, Tetrahedron: Asymmmetry<br />

1994, 5, 511–514; b) S. Gladiali, A. Dore, D. Fabbri, O. De Lucchi, G. Valle, J. Org. Chem.<br />

1994, 59, 6363–6371; c) S. Gladiali, D. Fabbri, Chem. Ber./ Rec. Trav. Chim. Pays Bas. 1997,<br />

130, 543–554; d) J. Durand, S. Gladiali, G. Erre, E. Zangrando, B. Milani, Organometallics<br />

2007, 26, 810–818.<br />

52 a) K. Tani, L. D. Brown, J. Ahmed, J. A. Ibers, M. Yokota, A. Nakamura, S. Otsuka, J. Am.<br />

Chem Soc. 1977, 99, 7876–7886; b) A. Cope, E. C. Friedrich, J. Am. Chem. Soc. 1968, 90,<br />

909–913.<br />

53 More recently the reaction was extended by Beller and co–workers applying representatives<br />

of ligand library 66. G. Erre, S. Enthaler, K. Junge, S. Gladiali, M. Beller, J. Mol. Catal. A,<br />

2008, 280, 148–155.<br />

54 F. Bitterer, O. Herd, M. Kühnel, O. Stelzer, N. Weferling, W. S. Sheldrick, J. Hahn, S. Nagel,<br />

N. Rösch, Inorg. Chem. 1998, 37, 6408–6417.<br />

35


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

55 G. Bringmann, A. J. P. Mortimer, P. A. Keller, M. J. Gresser, J. Garner, M. Breuning, Angew.<br />

Chem. 2005, 117, 5518–5563.<br />

56 T. Hayashi, K. Hayashizaki, T. Kiyoi, Y. Ito, J. Am. Chem. Soc. 1988, 110, 8153–8156; see<br />

also: L. Dahlenburg, V. Kurth, Inorg. Chim. Acta 2001, 319, 176–182; A. Terfort, H.<br />

Brunner, J. Chem. Soc. Perkin Trans. 1, 1996, 1467–1479.<br />

57 a) F. J. Jensen, M. Johannsen, Org. Lett. 2003, 5, 3025–3028; b) A. N. Cammidge, K. V. L.<br />

Crepy, Chem. Commun. 2000, 1723–1724; c) A. N. Cammidge, K. V. L. Crepy, Tetrahedron<br />

2004, 60, 4377–4386; d) A. N. Cammidge, K. V. L. Crepy, Chem. Comm. 2000, 1723–1724.<br />

58 Optically pure 2,2´–binaphthol is available on large scale from RCA (Reuter Chemische<br />

Apparatebau KG), Engesserstr. 4, 79108 Freiburg, Germany. (current price: 1 kg ~ 500 €)<br />

59 a) K. Junge, G. Oehme, A. Monsees, T. Riermeier, U. Dingerdissen, M. Beller, Tetrahedron<br />

Lett. 2002, 43, 4977–4980; b) K. Junge, G. Oehme, A. Monsees, T. Riermeier, U.<br />

Dingerdissen, M. Beller, J. Organomet. Chem. 2003, 675, 91–96; c) K. Junge, B. Hagemann,<br />

S. Enthaler, A. Spannenberg, M. Michalik, G. Oehme, A. Monsees, T. Riermeier, M. Beller,<br />

Tetrahedron: Asymmetry 2004, 15, 2621–2631; d) S. Enthaler, B. Hagemann, K. Junge, G.<br />

Erre, M. Beller, Eur. J. Org. Chem. 2006, 2912–2917.<br />

60 a) Y. Uozumi, M. Kawatsura, T. Hayashi, Org. Synth. 2002, 78, 1–13; b) D. Cai, J. F. Payack,<br />

D. R. Bender, D. L. Hughes, T. R. Verhoeven, P. J. Reider, Org. Synth., 1999, 76, 6–11.<br />

61 The synthesis of 2,2´–bistriflate-1,1´–binaphthyl (1.2 mol, 615 g, 92%) and 2,2´–dimethyl–<br />

1,1`–binaphthyl (0.74 mol, 201 g, 96%) was carried out in large scale by the group of Zhang<br />

(See reference 62).<br />

62 Aminophosphinite 70 was also used as ligand in the asymmetric hydrogenation of α–dehydro<br />

amino acid derivatives. The corresponding saturated compound was obtained with<br />

enantioselectivities up to 90% ee. By variation of the substituents on the amino functionality<br />

the enantioselectivity was improved to 96% ee (See reference: 56b).<br />

63 Y. Chi, X. Zhang, Tetrahedron Lett. 2002, 43, 4849–4852.<br />

64 a) W. Tang, W. Wang, Y. Chi, X. Zhang, Angew. Chem. Int. Ed. 2003, 42, 3506–3509; b) Y.<br />

Chi, Y.-G. Zhou, X. Zhang, J. Org. Chem. 2003, 68, 4120–4122; c) D. Xiao, X. Zhang,<br />

Angew. Chem. Int. Ed. 2001, 40, 3425–3428; d) D. Xiao, Z. Zhang, X. Zhang, Org. Lett.<br />

1999, 1, 1679–1681; Q. Dai, W. Yang, X. Zhang, Org. Lett. 2005, 7, 5343–5345; e) A. Lei,<br />

M. Chen, M. He, X. Zhang, Eur. J. Org. Chem. 2006, 4343–4347.<br />

65 H. Buschmann, H.-D. Scharf, N. Hoffmann, P. Esser, Angew. Chem. 1991, 103, 480–518.<br />

66 E. Alberico, I. Nieddu, R. Taras, S. Gladiali, Helv. Chim. Acta 2006, 89, 1716–1729.<br />

67 M. T. Reetz, G. Mehler, A. Meiswinkel, Tetrahedron: Asymmetry 2004, 15, 2165–2167.<br />

36


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

68 a) M. T. Reetz, X. Li, Angew. Chem. Int. Ed. 2005, 44, 2962–2964; b) A. Duursma, R. Hoen,<br />

J. Schuppan, R. Hulst, A. J. Minnaard, B. L. Feringa, Org. Lett. 2003, 5, 3111–3113; c) A.<br />

Duursma, D. Peña, A. J. Minnaard, B. L. Feringa, Tetrahedron: Asymmetry 2005, 16, 1901–<br />

1904; d) D. Peña, A. J. Minnaard, J. A. F. Boogers, A. H. M. de Vries, B. L. Feringa, Org.<br />

Biomol. Chem. 2003, 1, 1087–1089; e) C. Monti, C. Gennari, U. Piarulli, J. G. de Vries, A. H.<br />

M. de Vries, L. Lefort, Chem. Eur. J. 2005, 11, 6701–6717; f) K. Ding, H. Du, Y. Yuan, J.<br />

Long, Chem. Eur. J. 2004, 10, 2872–2884; g) J. G. de Vries, L. Lefort, Chem. Eur. J. 2006,<br />

12, 4722–4734.<br />

69 S. Enthaler, G. Erre, K. Junge, J. Holz, A. Börner, E. Alberico, I. Nieddu, S. Gladiali, M.<br />

Beller, Org. Process Res. Dev. 2007, 11, 568–577.<br />

70 a) W. D. Lubell, M. Kitamura, R. Noyori, Tetrahedron: Asymmetry 1991, 2, 543–554; b) X.-<br />

P. Hu, Z. Zheng, Org. Lett. 2005, 7, 419–422.<br />

71 a) B. D. Vineyard, W. S. Knowles, M. J. Sabacky, G. L. Bachman, D. J. Weinkauff, J. Am.<br />

Chem. Soc. 1977, 99, 5947–5952; b) K. Achiwa, T. Soga, Tetrahedron Lett. 1978, 13, 1119–<br />

1120; c) W. D. Lubell, M. Kitamura, R. Noyori, Tetrahedron: Asymmetry 1991, 2, 543–554;<br />

d) G. Zhu, Z. Chen, X. Zhang, J. Org. Chem. 1999, 64, 6907–6910; e) D. Heller, J. Holz, H.-<br />

J. Drexler, J. Lang, K. Drauz, H.-P. Krimmer, A. Börner, J. Org. Chem. 2001, 66, 6816–6817;<br />

f) M. Yasutake, I. D. Gridnev, N. Higashi, T. Imamoto, Org. Lett. 2001, 3, 1701–1704; g) D.<br />

Heller, H.-J. Drexler, J. You, W. Baumann, K. Drauz, H.-P. Krimmer, A. Börner, Chem. Eur.<br />

J. 2002, 8, 5196–5203; h) Y.-G. Zhou, W. Tang, W.-B. Wang, W. Li, X. Zhang, J. Am. Chem.<br />

Soc. 2002, 124, 4952–4953; i) W. Yang, S. Wu, X. Zhang, J. Am. Chem. Soc. 2003, 125,<br />

9570–9571; j) I. V. Komarov, A. Monsees, A. Spannenberg, W. Baumann, U. Schmidt, C.<br />

Fischer, A. Börner, Eur. J. Org. Chem. 2003, 138–150; k) J. You, H.-J. Drexler, S. Zhang, C.<br />

Fischer, D. Heller, Angew. Chem. 2003, 115, 942–945; m) J. Holz, A. Monsees, H. Jiao, J.<br />

You, I. V. Komarov, C. Fischer, K. Drauz, A. Börner, J. Org. Chem. 2003, 68, 1701–1707; n)<br />

J. Wu, X. Chen, R. Guo, C.-h. Yeung, A. S. C. Chan, J. Org. Chem. 2003, 68, 2490–2493; o)<br />

H.-P. Wu, G. Hoge, Org. Lett. 2004, 6, 3645–3647; p) Y. Hsiao, N. R. Rivera, T. Rosner, S.<br />

W. Krska, E. Njolito, F. Wang, Y. Sun, J. D. Armstrong III, E. J. J. Grabowski, R. D. Tillyer,<br />

F. Spindler, C. Malan, J. Am. Chem. Soc. 2004, 126, 9918–9919.<br />

72 S. Enthaler, G. Erre, K. Junge, D. Michalik, A. Spannenberg, S. Gladiali, M. Beller,<br />

Tetrahedron: Asymmetry, 2007, 18, 1288–1298.<br />

73 X. Jiang, M. van den Berg, A. J. Minnaard, B. Feringa, J. G. de Vries, Tetrahedron:<br />

Asymmetry 2004, 15, 2223–2229.<br />

74 P. Kasák, K. Mereiter, M. Widhalm, Tetrahedron: Asymmetry 2005, 16, 3416–3426.<br />

37


2. Phosphorous containing ligands in the field of asymmetric hydrogenation<br />

75 a) K. Junge, B. Hagemann, S. Enthaler, G. Oehme, M. Michalik, A. Monsees, T. Riermeier,<br />

U. Dingerdissen, M. Beller, Angew. Chem. 2004, 116, 5176–5179; Angew. Chem. Int. Ed.<br />

2004, 43, 5066–5069; b) B. Hagemann, K. Junge, S. Enthaler, M. Michalik, T. Riermeier, A.<br />

Monsees, M. Beller, Adv. Synth. Catal. 2005, 347, 1978–1986.<br />

76 G. Zanoni, S. Gladiali, A. Marchetti, P. Piccinini, I. Tredici, G. Vidari, Angew. Chem. Int. Ed.<br />

2004, 43, 846–849.<br />

77 L. Charruault, V. Michelet, R. Taras, S. Gladiali, J.-P. Genêt, Chem. Commun. 2004, 850–<br />

851.<br />

78 a) R. P. Wurz, G. C. Fu, J. Am. Chem. Soc. 2005, 127, 12234–12235; b) E. Vedejs, O.<br />

Daugulis, S. T. Diver, J. Org. Chem. 1996, 61, 430–431; c) J. E. Wilson, G. C. Fu, Angew.<br />

Chem. Int. Ed. 2006, 45, 1426–1429.<br />

38


3. Reduction of unsaturated compounds with homogeneous iron catalysts 39<br />

3. Reduction of unsaturated compounds with homogeneous iron<br />

catalysts<br />

Stephan Enthaler, Kathrin Junge, and Matthias Beller*, in B. Plietker (Ed.) “Iron catalysis in<br />

organic chemistry“ Wiley VCH, Weinheim, 2008.<br />

Contributions<br />

The subchapter “Hydrosilylation” was written by Dr. K. Junge.


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

Stephan Enthaler, Kathrin Junge, and Matthias Beller<br />

Leibniz-Institut für Katalyse e.V. an der <strong>Universität</strong> <strong>Rostock</strong>, Albert-Einstein-Str. 29a,<br />

18059 <strong>Rostock</strong>, Germany. Fax: 49-381-1281-5000; Tel: 49-381-1281-113; E-mail:<br />

matthias.beller@catalysis.de<br />

3.1. Introduction<br />

Within the different molecular transformations reduction processes plays one of the major<br />

roles in organic chemistry. Apart from stoichiometric reactions, transition metal catalyzed<br />

reactions offer an efficient and versatile strategy and present a key technology for the<br />

advancement of “green chemistry”, specifically for waste prevention, reducing energy<br />

consumption, achieving high atom efficiency and creating advantageous economics. 1 Among<br />

the most popular and extensive studied catalytic reactions with respect to industrial<br />

applications is the hydrogenation of unsaturated compounds containing C=C, C=O and C=N<br />

bonds since addition of molecular hydrogen provides high atom efficiency. 2 In this regard, for<br />

activation of the molecular hydrogen or hydrogen donors, transition metal catalysts are<br />

essential. Commonly expensive and rare late transition metals (Ir, Rh or Ru) are applied as<br />

catalyst core. Therefore, substitution by ubiquitous available, inexpensive and less toxic<br />

metals is one of the major challenges for future research. Consequently, the use of iron<br />

catalysts is especially desirable. 3 This chapter will summarize the impact of iron catalysts on<br />

hydrogenation and hydrosilylation chemistry and mainly directed to the reduction of C=C,<br />

C=O and C=N bonds. Further catalytic applications were summarized elsewhere e.g.<br />

reduction of nitro compounds or dehalogenations. 4<br />

3.2. Hydrogenation of carbonyl compounds<br />

The relevance of carbonyl hydrogenation processes is impressively emphasized by the broad<br />

scope of applications for example as building blocks and synthons for pharmaceuticals,<br />

agrochemicals, polymers, synthesis of natural compounds, auxiliaries, ligands and key<br />

intermediates in organic syntheses (Scheme 1). 5<br />

40


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

O<br />

N<br />

Orphenadrin (1)<br />

anticholinergic drug<br />

HO<br />

OH<br />

HCl<br />

H<br />

N OMe<br />

OMe<br />

(R)-Denopamine-Hydrochloride (2)<br />

β 1-receptor antagonist<br />

Scheme 1. Selected pharmaceuticals based on chiral alcohols. 6<br />

F<br />

OH<br />

N<br />

BMS 181100 (3)<br />

antipsychotica<br />

N<br />

N N<br />

In the past this field has been dominated by ruthenium, rhodium and iridium catalysts with<br />

extraordinary activities and furthermore superior enantioselectivities, however some<br />

investigations were carried out with iron catalysts. Early efforts were reported on the<br />

successful use of hydridocarbonyliron complexes (HFem(CO)n - ) as reducing reagent for α,β-<br />

unsaturated carbonyl compounds, dienes and C=N double bonds, albeit complexes were used<br />

in stoichiometric amounts. 7 The first catalytic approach was presented by Markó and coworkers<br />

on the reduction of acetone in the presence of Fe3(CO)12 or Fe(CO)5. 8 In this reaction<br />

the hydrogen is delivered by water under more drastic reaction conditions (100 bar, 100 °C).<br />

Addition of NEt3 as co-catalyst was necessary to obtain reasonable yields. The authors<br />

assumed a reaction of Fe(CO)5 with hydroxide ions to yield HFe(CO)4 - under liberation of<br />

carbon dioxide since basic conditions are present, and excluded formation of molecular<br />

hydrogen via water gas shift reaction. HFe(CO)4 - is believed to be the active catalyst, which<br />

transferred the hydride to the acceptor. The presented catalyst displayed activity in the<br />

reduction of several ketones and aldehydes (Scheme 2). 9<br />

R 1<br />

O<br />

R 2<br />

10 mol% Fe(CO) 5<br />

NEt 3<br />

R 1<br />

OH<br />

100 bar (H2:CO 98.5:1.5)<br />

4 150 °C, 3 h<br />

5<br />

R 2<br />

5a: R 1 = CH 3; R 2 = CH 3: conv.: >99%<br />

5b: R 1 = C 6H 5; R 2 = CH 3: conv.: 62%<br />

5c: R 1 = i-Bu; R 2 = CH 3: conv.: 30%<br />

5d: R 1 = C 6H 5; R 2 = H: conv.: 95%<br />

5e: R 1 = n-Pr; R 2 = H: conv.: 98%<br />

6: cyclohexanone: conv.: >99%<br />

Scheme 2. Reduction of ketones and aldehydes by using the Fe(CO)5/NEt3 system according<br />

to Markó et al.<br />

Later on, the group of Vancheesan referred on transfer hydrogenation of ketones utilizing 2propanol<br />

or 1-phenylethanol as hydrogen source (Scheme 3). 10 A phase transfer catalyst was<br />

essential to support the hydrogenation catalyst Fe3(CO)12 or Fe(CO)5 since a liquid-liquid<br />

biphasic system was used as solvent. Under mild reaction conditions several ketones were<br />

hydrogenated to the corresponding alcohols with turnover frequencies up to 13 h -1 .<br />

Mechanistic investigations indicated a similar process as reported by Markó et al. 8<br />

41<br />

F


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

R 1<br />

O<br />

R 2<br />

4 mol% Fe 3(CO) 12<br />

PTC: BzEt 3NCl or Aliquat 336<br />

NaOH<br />

R 1<br />

OH<br />

4<br />

water/2-PrOH<br />

or water/1-phenylethanol<br />

28 °C, 2.5-3 h<br />

5<br />

R 2<br />

5b: R 1 = C 6H 5; R 2 = CH 3: yield: 36%<br />

5f: R 1 = 4-Cl-C 6H 4; R 2 = CH 3: yield: 42%<br />

5g: R 1 = 4-Me-C 6H 4; R 2 = CH 3: yield: 28%<br />

5h: R 1 = Et; R 2 = CH 3: yield: 20%<br />

6: cyclohexanone: yield: 78%<br />

Schema 3. Phase transfer catalyzed transfer hydrogenation of ketones by Vancheesan et al.<br />

More recently, Gao et al. reported in a special chinese journal an asymmetric transfer<br />

hydrogenation (Scheme 4) based on [Et3NH][HFe3(CO)11] and chelating chiral ligands. 11 In<br />

the presence of enantiopure diaminodiphosphine iron catalysts they claim good conversion<br />

and enantioselectivity up to 98% ee (substrate 5j).<br />

NH HN<br />

PPh 2 Ph 2P<br />

(S,S)-6<br />

(R,R)-6<br />

R 1<br />

O<br />

R 2<br />

1 mol% [Et 3NH][HFe 3(CO) 11]<br />

ligand 6<br />

6 mol% KOH<br />

R 1<br />

2-PrOH<br />

4 45-82 °C, 3-21 h<br />

5<br />

OH<br />

*<br />

R 2<br />

ligand: (S,S)-6<br />

5b: R 1 = C 6H 5; R 2 = CH 3: 56% ee (R), yield: 92% (82 °C, 7 h)<br />

5i: R 1 = C 6H 5; R 2 =C(CH 3) 3: 93% ee (S), yield: 19% (82 °C, 7 h)<br />

5j: R 1 = CH(C 6H 5) 2; R 2 =CH 3: 98% ee (S), yield: 18% (65 °C, 21 h)<br />

5k: R 1 = C 6H 5; R 2 =Cy: 78% ee (S), yield: 73% (82 °C, 17 h)<br />

ligand: (R,R)-6<br />

5m: R 1 = C 6H 5; R 2 = C 2H 5: 72% ee (S), yield: 87% (45 °C, 4.5 h)<br />

5n: R 1 = C 6H 5; R 2 = CH(CH 3) 2: 78% ee (S), yield: 98% (45 °C, 3 h)<br />

Schema 4. Enantioselective transfer hydrogenation catalyzed by iron complexes by Gao et al.<br />

In 2006 we reported the application of an easy to adopt in situ concept composed of an iron<br />

source e.g. FeCl2 or Fe3(CO)12, monodentate phosphines and tridentate nitrogen ligands in the<br />

transfer hydrogenation of aliphatic and aromatic ketones using 2-propanol as hydrogen source<br />

(Scheme 5). 12 The influence of different reaction parameters on the reduction of model<br />

substrate acetophenone 5b were studied in detail. A crucial influence of base and base<br />

concentration was displayed since only sodium isopropoxide gave reasonable amount of<br />

product. In comparison to ruthenium-based transfer hydrogenations a higher temperature is<br />

necessary. Notably, the advantage of this in situ concept is underlined by easy tuneability of<br />

the iron catalyst, due to the broad availability of simple phosphines and amines. Plenty of<br />

phosphorous and amine ligands were subjected to the model reaction. However, best results<br />

were found with a catalyst composed of Fe3(CO)12, triphenylphosphine and 2,2´:6´,2´´terpyridine<br />

or, surprisingly, 3 equiv. of pyridine with respect to iron. The later one resembles<br />

42


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

a straightforward catalyst regarding to industrial applications. After several optimization steps<br />

an active catalyst was developed which exhibit comparable activity to Ru3(CO)12 based<br />

catalysts and was success<strong>full</strong> for the reduction of several ketones with good to excellent<br />

conversions. 13<br />

R 1<br />

O<br />

4<br />

R 2<br />

0.5 mol% FeCl 2<br />

0.5 mol% PPh 3/0.5 mol% TerPy<br />

25 mol% NaO-i-Pr<br />

2-PrOH, 100 °C, 7 h<br />

R 1<br />

OH<br />

5<br />

R 2<br />

5b: R 1 = C 6H 5; R 2 = CH 3: yield: 95%<br />

5f: R 1 = 4-Cl-C 6H 4; R 2 = CH 3: yield: 97%<br />

5g: R 1 = 4-Me-C 6H 4; R 2 = CH 3: yield: 83%<br />

5m: R 1 = C 6H 5; R 2 = C 2H 5: yield: 92%<br />

5o: R 1 = 4-MeO-C 6H 4; R 2 = CH 3: yield: 75%<br />

5p: R 1 = 2-MeO-C 6H 4; R 2 = CH 3: yield: >99%<br />

5q: R 1 = C 6H 5; R 2 = CH 2Cl: yield: 8%<br />

5r: R 1 = Cy; R 2 = CH 3: yield: 93%<br />

5s: R 1 = t-Bu; R 2 = CH 3: yield: >99%<br />

Scheme 5. In situ catalyst based on FeCl2/terpy/PPh3 in the transfer hydrogenation of ketones.<br />

Very recently the effectiveness of FeCl2/terpy/PAr3 catalysts is proven in the hydrogenation<br />

of α-substituted ketones, which are interesting 1,2-diol precursors (Scheme 6). 14 Excellent<br />

yields, chemoselectivities and activities (TOF: up to 2000 h -1 ) were achieved under optimized<br />

conditions.<br />

R 1<br />

O<br />

1 mol% FeCl2 1 mol% P(4-F-C6H4) 3/1 mol% TerPy<br />

OR<br />

50 mol% NaOH<br />

2<br />

R<br />

2-PrOH, 100 °C, 1 h<br />

1<br />

OH<br />

7 8<br />

OR 2<br />

43<br />

8a: R 1 = C 6H 5; R 2 = CH 3: yield: 92%<br />

8b: R 1 = C 6H 5; R 2 = C 6H 5: yield: 94%<br />

8c: R 1 = C 6H 5; R 2 = 4-Cl-C 6H 4: yield: >99%<br />

8d: R 1 = C 6H 5; R 2 = 2,6-(i-Pr) 2C 6H 4: yield: 45%<br />

8e: R 1 = 4-Me-C 6H 5; R 2 = C 6H 5: yield: 99%<br />

8f: R 1 = 4-MeO-C 6H 5; R 2 = C 6H 5: yield: 85%<br />

8g: R 1 = t-Bu; R 2 = C 6H 5: yield: 22%<br />

Scheme 6. Reduction of α-substituted ketones in the presence of iron catalyst.<br />

Apart from synthetic aspects we attained some useful informations concerning the<br />

mechanism, even if the “real” catalyst structure is so far unclear. Various experiments proved<br />

the presence of a homogeneous catalyst. Following the original works on Ru-, Rh- or Ircatalyst,<br />

deuterated hydrogen donors were applied since for transfer hydrogenation two<br />

common mechanisms are established, nominated as direct hydrogen transfer via formation of<br />

a six-membered cyclic transition state constituted of metal, hydrogen-donor and -acceptor,<br />

and secondly the hydridic route, which is subdivided into two pathways, the monohydride and<br />

dihydride mechanism (Scheme 7). More specifically, the formation of monohydride-metalcomplexes<br />

promote an exclusive hydride transfer from carbon (donor) to carbonyl carbon<br />

(acceptor), whereas a hydride transfer via dihydride-metal-complexes leads to no accurate


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

prediction of hydride resting state, for the reason that the former hydride was transferred to<br />

carbonyl carbon (acceptor) as well as to the carbonyl oxygen (acceptor). 15<br />

O<br />

[M-H*]<br />

H<br />

O<br />

*H<br />

monohydride mechanism dihydride mechanism<br />

-acetone<br />

OH<br />

H*<br />

O<br />

[H-M-H*]<br />

H<br />

O<br />

*H<br />

-acetone<br />

OH<br />

H*<br />

44<br />

OH*<br />

H<br />

50% 50%<br />

Scheme 7. Hydridic route established for transfer hydrogenation with Rh-, Ru- and Ircatalysts.<br />

To specify the position and the nature of the transferred hydride, the reaction was performed<br />

with 2-propanol-d1 as solvent/donor, sodium 2-propylate as base and Fe3(CO)12/PPh3/TerPy<br />

as catalyst under optimized conditions. In the transfer hydrogenation of acetophenone a<br />

mixture of two deuterated 1-phenylethanols was obtained (Scheme 8, 9a and 9b). The ratio<br />

between 9a and 9b (85:15) indicated a specific migration of the hydride, albeit some<br />

scrambling was detected. However, the incorporation is in agreement with the monohydride<br />

mechanism, implying a formation of metal monohydride species in the catalytic cycle.<br />

O<br />

5b<br />

[Fe]/NaO-i-Pr<br />

2-PrOD, 100 °C<br />

O<br />

H<br />

D<br />

+<br />

O<br />

D<br />

H<br />

9a: 85% 9b:15%<br />

Scheme 8. Deuterium incorporation catalyzed by Fe3(CO)12/terpy/PPh3 system under transfer<br />

hydrogenation conditions.<br />

Parallel to the work on in situ three component catalysts we developed a nature inspired in<br />

situ catalyst based on iron porphyrin complexes since a high stability of the complexes are<br />

known and inertness against oxygen and moisture is feasible. 16 The porphyrin catalysts<br />

achieved even higher activities than three component system in the transfer hydrogenation of<br />

ketones (Scheme 9). After optimization of reaction parameters turnover frequencies up to 642<br />

h -1 at low catalyst loadings (0.01 mol%) were attained. A ligand screening emphasized<br />

porphyrin 10 as module of a highly active catalyst and makes it suitable for further<br />

investigations on substrate variation. Various ketones were reduced in good to excellent<br />

yields. Notably, also naturally occurring hemin has been used in this study. Even if lower


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

activity was gained it is a worthwhile system, due to the abundance and easy handling.<br />

Advantageously, no pre-catalyst formation is necessary.<br />

N<br />

N<br />

N<br />

NH N<br />

N<br />

HN<br />

10<br />

N<br />

R 1<br />

O<br />

R 2<br />

0.17 mol% Fe 3(CO) 12<br />

0.5 mol% 10<br />

8.5 mol% NaO-i-Pr<br />

2-PrOH<br />

100 °C, 7 h<br />

R 1<br />

OH<br />

5f: R 1 = 4-Cl-C 6H 4; R 2 = CH 3: yield: 95%<br />

5b: R 1 = C 6H 5; R 2 = CH 3: yield: 94%<br />

5g: R 1 = 4-Me-C 6H 4; R 2 = CH 3: yield: 68%<br />

5o: R 1 = 4-MeO-C 6H 4; R 2 = CH 3: yield: 72%<br />

5p: R 1 = 2-MeO-C 6H 4; R 2 = CH 3: yield: >99%<br />

5m: R 1 = C 6H 5; R 2 = C 2H 5: yield: 87%<br />

5r: R 1 = Cy; R 2 = CH 3: yield: 89%<br />

5s: R 1 = t-Bu; R 2 = CH 3: yield: 90%<br />

Scheme 9. Biomimetic transfer hydrogenation of ketones with iron porphyrin catalysts.<br />

Applying Fe porphyrin catalysts and 2-propanol as hydrogen donor various α-hydroxyl-<br />

protected ketones are reduced to the corresponding mono protected 1,2-diols in good to<br />

excellent yield after optimization (Scheme 10). 17 Remarkable, addition of small amounts of<br />

water (5–10 mol%) boosted the catalysts activity up to 2500 h -1 at low catalyst loadings (0.01<br />

mol%).<br />

Cl<br />

N<br />

Cl<br />

NH N<br />

Cl<br />

HN<br />

11<br />

Cl<br />

R 1<br />

O<br />

1 mol% FeCl 2<br />

1 mol% 11<br />

50 mol% NaOH<br />

2-PrOH, 100 °C, 2 h<br />

8a: R 1 = C 6H 5; R 2 = CH 3: yield: >99%<br />

8b: R 1 = C 6H 5; R 2 = C 6H 5: yield: 92%<br />

8c: R 1 = C 6H 5; R 2 = 4-Cl-C 6H 4: yield: >99%<br />

8d: R 1 = C 6H 5; R 2 = 2,6-(i-Pr) 2C 6H 4: yield: 74%<br />

8e: R 1 = 4-Me-C 6H 5; R 2 = C 6H 5: yield: >99%<br />

8f: R 1 = 4-MeO-C 6H 5; R 2 = C 6H 5: yield: 83%<br />

8g: R 1 = t-Bu; R 2 = C 6H 5: yield: 48%<br />

Scheme 10. Reduction of α-substituted ketones in the presence of iron porphyrin catalysts.<br />

OR 2<br />

3.3. Hydrogenation of Carbon-Carbon Double Bounds<br />

Since the 60 th of the last century tremendous efforts were undertaken in the field of<br />

homogeneous hydrogenation of C-C double bonds emphasized by innumerable reports and<br />

later on rewarded by integrations in several industrial processes (Scheme 11). However, an<br />

R 1<br />

OH<br />

R 2<br />

45<br />

OR 2


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

even broader use in industry is expected by the availability of similar active and selective<br />

hydrogenation catalysts containing inexpensive metals (e.g. Fe and Cu).<br />

HO<br />

OH<br />

O<br />

NH 2<br />

OH<br />

L-DOPA (12)<br />

anti Parkinson´s disease<br />

O<br />

SH<br />

N<br />

H COOH<br />

N-Acetylcystein (13)<br />

mucolytic therapy<br />

H 2N<br />

COOH<br />

H<br />

N COOMe<br />

O<br />

Aspartam (14)<br />

artificial sweetener<br />

O N<br />

O<br />

O<br />

Cl<br />

F<br />

46<br />

Flamprop-isopropyl (15)<br />

herbizide<br />

Scheme 11. Selected pharmaceuticals, which are accessible by asymmetric C-C double bond<br />

hydrogenation. 5,18<br />

Indeed iron catalysts are known to transfer hydrogen to C-C double bonds. Early examples<br />

were carried out with metal salts activated by aluminium compounds in the hydrogenation of<br />

non-funtionalized olefins at comparable mild reaction conditions (0–35 °C and low hydrogen<br />

pressure). 19 Some research groups focused on the application of iron carbonyls due to the<br />

easier removal of the carbonyl ligands, for generating an active site, compared to halogens<br />

and avoidance of activating reagents. 20 Unsaturated carboxylic acid derivatives, for instance<br />

methyl linoleate and methyl linolenate, which contain two or three non-conjugated double<br />

bonds, respectively, were described in the mid of the 1960s (Scheme 12). 21,22 As catalyst<br />

precursor Fe(CO)5 was utilized at high temperature. The authors proposed as first step an<br />

isomerization of the double bonds catalyzed by Fe(CO)5 to obtain various dienes or trienes.<br />

Hydrogenation process was observed in case of conjugated dienes to yield monoenes via an<br />

iron-carbonyl-diene complex 18, while dienes were formed for the hydrogenation of methyl<br />

linolenate 16. The activity of the system was improved, when the iron-carbonyl-dienecomplex<br />

18 is used as precursor instead of Fe(CO)5. Analysis of the reaction mixture<br />

displayed mononenes as major products with unselective double bond distribution<br />

accompanied by small amounts of <strong>full</strong>y saturated compounds. Later on, Cais and co-worker<br />

studied the reaction of methyl hexa-2,4-dienoate as a model for fatty acids in more detail and<br />

reported several mechanistic considerations. 23


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

(CH 2) 4CH 3 (CH 2) 7COOMe<br />

H 3C(H 2C) x<br />

16<br />

Fe(CO) 5<br />

(CH 2) yCOOMe<br />

10 mol% Fe(CO) 5<br />

cyclohexane, H 2 (28 bar), 180 °C, 7 h<br />

Fe(CO) 5<br />

H 3C(H 2C) 4<br />

H3C(H2C) x Fe<br />

(CH2) yCOOMe<br />

OC CO<br />

CO 18<br />

Scheme 12. Fe-catalyzed hydrogenation of methyl linoleate.<br />

17<br />

H 3C(H 2C) v<br />

47<br />

(CH 2) 7COOMe<br />

(CH 2) wCOOMe<br />

Tajima and Kunioka described the hydrogenation of conjugated diolefins, in particular 1,3butadiene<br />

in the presence of activated [CpFe(CO)2Cl]. 24 AlEt3 and PhMgBr were tested as<br />

activation reagents, thereby the AlEt3-activated complex showed the best performance. The<br />

authors proposed the formation of iron-alkyl species which undergoes hydrogenolysis to<br />

create an active metal-hydride. The composition of the final mixture is dominated by cis-but-<br />

2-ene and trans-but-2-ene (ratio ~1:1). Furthermore traces of 1-butene were observed,<br />

whereas the catalyst was completely inactive for the hydrogenation of the mono-olefin to<br />

yield n-butane.<br />

19<br />

2.5 mol% CpFe(CO) 2Cl<br />

6 mol% AlEt 3 or 8.8 mol% PhMgBr<br />

benzene, H 2, 40-45 °C, 6 h<br />

20a<br />

+ +<br />

20b<br />

AlEt 3 conversion: 99%<br />

20a: 2%; 20b: 46%;20c: 52%<br />

PhMgBr conversion: 85%<br />

20a: 1%; 20b: 38%; 20c: 37%<br />

Scheme 13. Hydrogenation of butadiene with CpFe(CO)2Cl (Cp = cyclopentadienyl).<br />

Nishiguchi and Fukuzumi reported on the transfer hydrogenation of 1,5-cyclooctadiene (cod)<br />

in the presence of catalytic amounts of FeCl2(PPh3)2 (10 mol%) at high temperature (up to<br />

240 °C). 25 As hydrogen source polyhydroxybenzenes, for instance pyrogallol, pyrocatechol or<br />

hydrochinone, have been proven to be superior for this reaction, while primary and secondary<br />

alcohols led to lower conversion. The hydrogenation of 1,5-cyclooctadiene led to cyclooctane<br />

20c


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

as main-product, but the unsaturated cyclooctene and isomerization products were also<br />

observed.<br />

Later on Wrighton and co-workers recognized a reactivity increase with respect to<br />

hydrogenation rate when Fe(CO)5 was treated with light. 26,27 The authors assumed an<br />

expulsion of one carbonyl ligand to form an unsaturated species induced by near-ultraviolet<br />

irradiation. After complexation of the olefin again activation by light removed another<br />

carbonyl ligand to allow hydrogen coordination/activation. A number of unsubstituted olefins<br />

(e.g. ethylene, propylene, cis-3-hexene, cyclopentene, cyclooctene), were hydrogenated to<br />

give alkanes in moderate yield (up to 50%), while the catalyst is inactive for sterical<br />

demanding olefins (e.g. 1,2-dimethylcyclohexene) or aldehydes and nitriles. Furthermore,<br />

acetylenes were subjected to this photo-catalyzed reduction. However, only low yield (


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

At the beginning of the 90 th of the last century a switch to iron catalysts stabilized by either<br />

phosphines or nitrogen ligands took place. Here, Bianchini and co-workers carried out a<br />

detailed comparative study on the transfer hydrogenation of α,β-unsaturated ketones by<br />

nonclassical trihydride iron, ruthenium and osmium complexes containing tetradentate<br />

phosphines (Scheme 15). 30,31 In the presence of cyclopentanol as hydrogen donor several α,β-<br />

unsaturated ketones were hydrogenated to the corresponding saturated ketones with good to<br />

excellent selectivity under mild reaction conditions (Scheme 15). In some cases, e.g.<br />

benzylideneacetone or 2,3-cyclohexenone, the unsaturated or saturated alcohols were formed<br />

by the catalyst 26, thereby also good selectivity was noticed. Noteworthy, no co-catalyst, e.g.<br />

base, was necessary to activate the catalyst or the hydrogen donor, which is needed for other<br />

catalyst systems. However, for carbonyl hydrogenation, e.g. acetophenone and<br />

cyclohexanone, the iron catalyst displayed only low activity. No activity was found for<br />

aldehydes and C=C systems without additional carbonyl functionality.<br />

P<br />

P1 P1 Fe<br />

P1 P 1 = PPh 2<br />

R R1 dioxane,<br />

R1 H<br />

H<br />

H<br />

26<br />

O<br />

1 mol% 26<br />

cyclopentanol<br />

80 °C, 1-7 h<br />

O<br />

OH<br />

OH<br />

R<br />

+<br />

R R1 +<br />

R R1 A B C<br />

27a: R = C 6H 5; R 1 = CH 3: conv.: 95%, B: 95%<br />

27b: R = C 6H 5; R 1 = C 6H 5: conv.: 30%, A: 30%<br />

27c: R = CH 3; R 1 = C 6H 5: conv.: 7%, A: 7%<br />

27d: R = CH 3; R 1 = C 2H 5: conv.: 19%, A: 19%<br />

27e: R = CH 3; R 1 = CH 3: conv.: 100%, A: 100%<br />

28: 2-cyclohexene: conv.: 72%, B: 44%, C: 28%<br />

29: 2-cyclopentene: conv.: 100%, A: 84%, B: 16%<br />

Scheme 15. Application of nonclassical iron trihydride complex in transfer hydrogenation<br />

according to Bianchini and co-workers.<br />

Reduction of styrene was reported by Kano et al. using NaBH4 as hydrogen source in the<br />

presence of iron porphyrin complex 30 (Scheme 16). 32 Good turnover frequencies up to 81 h -1<br />

and good chemoselectvities are obtained in protic solvents with catalyst loading of 1 mol% at<br />

room temperature. The reaction has been proven to be a radical process since to some extent<br />

2,3-diphenylbutane was detected. A similar approach was carried out by Sakaki and co-<br />

workers on the hydrogenation of α,β-unsaturated esters, who reported improvement of the<br />

catalyst activity (turnover frequency, tof = 4580 h -1 ). 33 A detailed study of the reaction<br />

mechanism displayed a crucial influence of the protic solvent because the hydride is assigned<br />

by NaBH4 and the proton by methanol.<br />

49


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

N N<br />

Fe<br />

N N<br />

+<br />

Cl -<br />

30<br />

O<br />

OEt<br />

0.02 mol% 30<br />

200 mol% NaBH4 THF-MeOH (9:1)<br />

O<br />

OEt<br />

31<br />

30 °C, 1 h<br />

32<br />

TOF = 4580 h -1<br />

Scheme 16. Reduction of α,β-unsaturated ester catalyzed by iron porphyrin 30.<br />

More recently, the group of Chirik have shown elegantly the application of low valent iron<br />

complexes in the hydrogenation of various C-C double and C-C triple bounds. 34 Based on the<br />

mentioned work of Wrighton et al. an approach to stabilized 14 electron L3Fe(0) fragments<br />

was presented. The catalyst precursors were synthesized by reduction of dihalogen complexes<br />

33 containing a tridentate pyridinediimine ligand with sodium amalgam or with sodium<br />

triethylborohydride under an atmosphere of nitrogen (Scheme 17). The obtained bisdinitrogen<br />

complexes 34 are relative labile compounds and a loss of one equivalent of<br />

dinitrogen in solution at room temperature or an easy exchange against hydrogen or alkynes<br />

occurred. The catalyst activity was studied in the field of C-C double and C-C triple bond<br />

hydrogenation. Applying 0.3 mol% of the iron catalyst simple olefins such as 1-hexene or<br />

cyclohexene were hydrogenated with high turnover frequencies up to 1814 mol/h under<br />

comparatively mild reaction conditions (4 atm of hydrogen pressure and room temperature).<br />

The reaction was also carried out under preferred solvent-free conditions in neat substrates<br />

and leading to comparable activity. The scope and limitation of the catalyst system was<br />

demonstrated in the hydrogenation of various olefin units enclosing geminal, internal and<br />

trisubstituted olefins and furthermore diolefins. In the course of this it was ascertained that<br />

best activity was attained for terminal olefins > internal olefins = geminal olefins ><br />

trisubstituted olefins. In addition, one example of a functionalized olefin was mentioned.<br />

However, only a diminished activity was observed for dimethyl itaconate, albeit using higher<br />

catalyst loadings. In the case of cyclohexene some detailed investigations emphasized a<br />

deactivation of the catalyst by arene complexation either from the solvent or the aryl groups in<br />

the ligand. 35 Noteworthy, when comparing activity of catalyst 34 (1814 mol/h) with common<br />

catalysts e.g. Pd/C (TOF = 366 mol/h), RhCl(PPh3)3 (10 mol/h) or [Ir(cod)(PCy3)(py)]PF6 (75<br />

mol/h) in the hydrogenation of 1-hexene under optimized conditions a significant higher value<br />

was reached with the iron catalyst.<br />

50


R 1<br />

35<br />

R 2<br />

3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

Ar<br />

N<br />

N<br />

Fe<br />

N<br />

Ar<br />

0.3 mol% 34<br />

H 2 (4 atm)<br />

X X<br />

33a = Cl<br />

33b = Br<br />

toluene, 22 °C, 2 h<br />

R 1<br />

36<br />

method A<br />

N 2 (1 atm), Na(Hg), pentane<br />

method B<br />

N 2(1 atm), NaBEt 3H, toluene<br />

R 2<br />

Ar<br />

N<br />

N<br />

Fe<br />

N<br />

Ar<br />

N 2<br />

36a: R 1 = C 6H 5; R 2 = H: TOF = 1344 mol/h<br />

36b: R 1 = C 6H 5; R 2 =CH 3: TOF = 104 mol/h<br />

36c: R 1 = n-Bu; R 2 = H: TOF = 1814 mol/h<br />

36d: R 1 =CH 2=CH(CH 2) 2; R 2 = H: TOF = 363 mol/h<br />

36e: R 1 = -COOCH 3; R 2 = CH 2COOMe: TOF = 3 mol/h<br />

37: cylohexene: TOF = 57 mol/h<br />

34<br />

N 2<br />

Ar = 2,6-( i Pr) 2C 6H 3<br />

Scheme 17. Preparation of catalyst precursors containing tridentate nitrogen ligands and<br />

abilities in catalytic reaction according to Chirik and co-workers.<br />

Further on the group of Chirik studied the influence of replacing the imino functionalities in<br />

ligand system 34 by phosphino groups (Scheme 18). 36 A different coordination mode was<br />

found since only one nitrogen ligand was replaced by hydrogen. The unstable complex 38 was<br />

also tested in the hydrogenation of 1-hexene, but no improvement of activity was observed.<br />

N<br />

H<br />

P i Pr 2<br />

Fe<br />

N 2<br />

P i Pr 2<br />

H<br />

38<br />

35c<br />

0.3 mol% 37<br />

H 2 (4 atm)<br />

pentane, 23 °C, 3 h<br />

Scheme 18. Application of iron aminodiphosphine complexes in hydrogenation reaction.<br />

36c<br />

51<br />

conv.: 98%<br />

Using the same synthetic method as described for catalyst 34 also diime complexes 39 were<br />

attainable (Scheme 19). 37 Due to the instability of dinitrogen complexes stabilization was<br />

carried out by introducing more suitable ligands e.g. alkynes, olefins or diolefins. However, in<br />

the hydrogenation of 1-hexene under comparable conditions only low activities were<br />

observed. Apart from synthesis and application of low valent iron complexes Chirik et al.<br />

carried out some mechanistic investigations. The suggestion of the catalytic cycle is presented<br />

in Scheme 20. Initially, an unsaturated iron complex is formed by expulsion of both<br />

dinitrogen molecules.


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

Ar<br />

N<br />

Fe<br />

N<br />

Ar<br />

L<br />

39<br />

35c<br />

0.3 mol% 39<br />

H 2 (4 atm)<br />

pentane, 22 °C, 3 h<br />

39a: Ar = 2,6-( i Pr) 2C 6H 3; L = (Me 3Si) 2C 2 TOF = 4 mol/h<br />

39b: Ar = 2,6-( i Pr) 2C 6H 3; L = cod TOF = 90 mol/h<br />

39c: Ar = 2,6-( i Pr) 2C 6H 3; L = coe TOF = 90 mol/h<br />

Scheme 19. Hydrogenation with low valent iron complexes. (cod = 1,5-cyclooctadiene, coe =<br />

cyclooctene).<br />

Next coordination of olefin takes place, which is preferred since also activation of hydrogen is<br />

feasible. After olefin coordination oxidative addition of hydrogen yields a formally 18<br />

electron complex. Insertion of the olefin obtained an alkyl complex, which recreate the<br />

starting complex via reductive elimination. Notably, the olefin complex also supports an<br />

isomerization of the double bond; hence an extension of possible intermediates is conceivable.<br />

H<br />

[Fe]<br />

R<br />

R<br />

[Fe](N 2) 2<br />

[Fe]<br />

-2N 2<br />

Scheme 20. Mechanism for the hydrogenation process proposed by Chirik et al.<br />

[Fe]<br />

3.4. Hydrogenation of imines and similar compounds<br />

H<br />

R<br />

Until today only a scarce number of hydrogenations of unsaturated CN functionalities have<br />

been described utilizing iron catalysts. On the one hand hydrogenation of N-<br />

H<br />

[Fe]<br />

R<br />

H 2<br />

R<br />

36c<br />

52


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

benzylideneaniline to yield N-benzylaniline was carried out by Markó et al. in presence of<br />

catalytic amounts of [NEt3H][HFe(CO)4], which was also applied for ketones/aldehydes<br />

reduction (vide supra). 38 Even if good to excellent activities were obtained in various<br />

solvents, for proceeding the reaction harsh conditions are required (150 °C, 100 bar). On the<br />

other hand Kaesz and co-workers reported the reduction of nitrogen heterocycles under water-<br />

gas shift conditions. 39 As catalyst Fe(CO)5 was used under high temperature and pressure.<br />

Moderate turnover numbers in the range of 18–37 were obtained. In the case of isoquinoline<br />

42 as main product the formylated compound 43 is observed.<br />

N<br />

1.8 mol% Fe(CO) 5, KOH<br />

H 2O, MeOH<br />

CO (500 psi), 150 °C, 15 h<br />

N<br />

H<br />

N<br />

1.8 mol% Fe(CO) 5, KOH<br />

H 2O, MeOH<br />

CO (500 psi), 150 °C, 15 h<br />

40 41 42 43<br />

Scheme 21. Hydrogenation of aromatic nitrogen heterocycles.<br />

3.5. Catalytic hydrosilylations<br />

In addition to catalytic reductions with molecular hydrogen or hydrogen donors also silanes<br />

represent useful reducing agents. 40<br />

Most examples in literature for hydrosilylation with iron complexes as catalyst are known for<br />

Fe(CO)5 or related iron carbonyl compounds. 41 The first use of iron pentacarbonyl was<br />

reported for the reaction of silicon hydrides with olefins at 100–140 °C to form saturated and<br />

unsaturated silanes according to Scheme 22. 42,43<br />

R 3SiH + RCH=CH 2<br />

R = Cl, OC 2H 5, C 2H 5<br />

Fe(CO) 5<br />

neat, 100-140 °C<br />

Scheme 22. Fe-catalyzed hydrosilylation of C-C double bonds.<br />

R 3SiCH 2CH 2R<br />

44<br />

R 3SiCH=CHR + H 2<br />

An excess of olefin favours the formation of the unsaturated product 44 while the silylated<br />

compound 45 dominated at higher ratio of silane to alkene. The described reaction was also<br />

carried out in the presence of colloidal iron and lead to results similar to those obtained with<br />

Fe(CO)5. Hydrosilylation of vinyltrimethylsilane 46 with chlorosilanes R3SiH using the<br />

modified iron carbonyl complex (CH2=CHSiMe3)Fe(CO)4 44,45 gave a mixture of α- and β-<br />

45<br />

N<br />

53<br />

O


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

isomers of the silylated product 47 and the unsaturated product 48 attained by<br />

dehydrogenative silylation (Scheme 23).<br />

R 3SiH + Me 3SiCH=CH 2<br />

46<br />

2.5-5 mol%<br />

(CH 2=CHSiMe 3)Fe(CO) 4<br />

neat, 70-80 °C, 3-4 h<br />

Scheme 23. Hydrosilylation of vinyltrimethylsilane 46.<br />

CH 3<br />

Me3SiCHSiR3 α-47<br />

+ Me3SiCH2CH2SiR3 β-47<br />

Me 3SiCH=CHSiR 3 + EtSiMe 3<br />

The ratio of formed compounds α-47, β-47 and 48 depends on the nature of the used<br />

hydrosilane R3SiH. There exists an indirect relationship between the yield of the unsaturated<br />

silane 48 and the catalytic activity of R3SiH (Cl3SiH > Cl2MeSiH > ClMe2SiH > Me3SiH).<br />

With exception of Me3SiH the α-isomer is mainly formed in all studied reactions. Iron<br />

pentacarbonyl is known to hydrosilylate functionalized olefins (Scheme 24). 46 For instance<br />

hydrosilylation of acrolein 49 with Et3SiH gave up to 95% of a mixture of cis- and trans-<br />

MeCH=CHOSiEt3 50 while acrolein diethyl acetal 51 forms the Markovnikov product<br />

Et3SiCH2CH2CH(OEt)2 52 and Et3SiOEt via C-O bond cleavage. Allylic alcohol reacts in the<br />

presence of Fe(CO)5 to give CH2=CHCH2OSiEt3 54 and propylOSiEt3 55.<br />

CH2=CHCHO + Et3SiH 49<br />

CH 2=CHCH(OEt) 2 + Et 3SiH<br />

51<br />

CH 2=CHCH 2OH + Et 3SiH<br />

53<br />

1-4 mol% Fe(CO) 5<br />

neat, 140 °C, 4 h<br />

1 mol% Fe(CO) 5<br />

neat, 140 °C, 4 h<br />

1 mol% Fe(CO) 5<br />

neat, 140 °C, 4 h<br />

Scheme 24. Hydrosilylation of functionalized olefins.<br />

48<br />

CH 3CH=CHOSiEt 3<br />

50<br />

Et3SiCH2CH2CH(OEt) 2<br />

52<br />

CH2=CHCH2OSiEt3 + CH3(CH2) 2OSiEt3 54<br />

55<br />

In 1993 the group of Murai has examined the effectiveness of the iron-triad carbonyl<br />

complexes, Fe(CO)5, Fe2(CO)9, Fe3(CO)12, as catalysts for the reaction of styrene with<br />

triethylsilane. 47 While Fe(CO)5 showed no catalytic activity Fe2(CO)9, and Fe3(CO)12 form<br />

selectively β-silylstyrene and ethylbenzene. Interestingly, Fe3(CO)12 is the catalyst that<br />

exhibited the highest selectivity. This trinuclear iron carbonyl catalyst was also successful<br />

54


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

applied for the reaction of different p-substituted styrenes with Et3SiH giving only the (E)-β-<br />

triethylstyrenes in 66–70% yield.<br />

R<br />

56<br />

+ Et3SiH 0.3 or 0.7 mol%<br />

Fe2(CO) 9 or Fe3(CO) 12<br />

SiEt3 +<br />

benzene, 40-80 °C,<br />

24-72 h<br />

R R<br />

57a: R = H; yield: 36-89%<br />

57b: R = CH3; yield: 67%<br />

57c: R = Cl; yield: 66%<br />

57a: R = OCH3; yield: 70%<br />

Scheme 25. Hydrosilylation of styrene derivatives.<br />

Hydrosilylation reactions catalyzed by iron carbonyl compounds often occur under drastic<br />

thermal conditions. Wrighton et al. have reported a photocatalyzed reaction of trialkylsilanes<br />

with alkenes in the presence of Fe(CO)5 at low temperature (0–50 °C). 48 It is well known that<br />

irradiation of mononuclear metal carbonyls leads to efficient dissociative loss of CO to yield<br />

coordinative unsaturated, 16-electron, intermediates (Scheme 26). 26<br />

Fe(CO) 5<br />

H<br />

R3Si H<br />

R3Si + HSiR 3<br />

hv<br />

Fe(CO) 4<br />

Fe(CO) 3<br />

-CO<br />

+ CO<br />

+<br />

Fe(CO) 4<br />

- HSiR 3<br />

- + HSiR 3<br />

Fe(CO) 4<br />

hv hv<br />

H<br />

R3Si Fe(CO) 3<br />

+<br />

Fe(CO) 3 + CO<br />

Scheme 26. Formation of the catalytic active species by irradiation.<br />

Irradiation of Fe(CO)5 in the presence of trialkylsilane R3SiH and alkene initially yields a<br />

mixture of Fe(CO)4(alkene) and (H)(SiR3)Fe(CO)4 which were determined by infrared<br />

spectroscopy. (H)(SiR3)Fe(CO)3(alkene) is postulated to be the catalytic active species<br />

58<br />

55


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

photogenerated from these intermediates according to Scheme 26. Continuous irradiation is<br />

needed to maintain a steady-state concentration of the repeating unit “Fe(CO)3”. Such an iron<br />

tricarbonyl complex was also detected by infrared spectroscopy in the photoinduced addition<br />

of R3SiH to 1,3-butadiene with Fe(CO)5. 49 Polymer-anchored ironcarbonyl species were<br />

described to catalyze the hydrosilylation of 1-pentene with HSiEt3 to give a mixture of pentyl-<br />

and pentenylsilanes. 50 The different pathways of the mechanism of transition metal-catalyzed<br />

hydrosilylation are summarized in Scheme 27. A commonly proposed mechanism involves<br />

the insertion of the olefin into the Fe-H bond of complex (H)(SiR3)Fe(CO)3(alkene) followed<br />

by reductive elimination of the alkyl and the silyl group to form an alkylsilane (pathway B).<br />

In an alternative mechanism insertion of the olefin into the Fe-Si bond of complex<br />

(H)(SiR3)Fe(CO)3(alkene) is discussed (pathway A). This route which is also suggested for<br />

the use of Fe3(CO)12 explains the formation of vinylsilanes as by-product in<br />

ydrosilylations. 51<br />

h<br />

H<br />

H<br />

H<br />

Fe(CO) 3<br />

+ HSiR 3<br />

C<br />

SiR 3<br />

Fe(CO) 3<br />

SiR 3<br />

SiR 3<br />

+<br />

H<br />

H<br />

A<br />

R 3Si<br />

Fe(CO) 3<br />

H<br />

R 3Si<br />

+ HSiR 3<br />

H<br />

Fe(CO) 3<br />

Fe(CO) 3<br />

B<br />

+ HSiR 3<br />

R 3Si<br />

R 3Si<br />

Fe(CO) 3<br />

Scheme 27. Proposed catalytic cycle for the hydrosilylation of olefins in the presence of<br />

Fe(CO)5.<br />

56


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

The photoinduced alkene insertion into the Fe-Si bond of (η 5 -C5Me5)Fe(CO)R was shown to<br />

be reversible (reaction C). 52 A reactivity study of complex (H)Fe(CO)4SiPh3 with<br />

nucleophiles allowed a better understanding of the underlying reaction processes. 53 Under<br />

photochemical conditions the activation or the substitution of the carbonyl ligands supports<br />

the classical mechanism involving insertion of the coordinated olefin in the Fe-Si (A) or the<br />

Fe-H (B) bonds. Under thermal conditions, however, a direct addition of the Fe-H group by a<br />

radical or ionic process can occur as observed in the reaction of (H)Fe(CO)4SiPh3 with<br />

isoprene. These mechanistic investigations for photo-generated hydrosilylation have been<br />

transmitted to the reaction of the iron carbonyl complex (CH2=CHSiMe3)Fe(CO)4 with<br />

vinylsilane and R3SiH. 45 Tetrahedral heterometallic clusters containing iron have proven to be<br />

suitable catalysts in the photoinitiated hydrosilylation of acetophenone with triethylsilane. 54<br />

Although a chiral FeCoMoS tetrahedron was used only racemic product and racemic cluster<br />

have been isolated. This fact can be explained by photo-racemization of the chiral catalyst<br />

which proceeds faster than hydrosilylation reaction. Instead of irradiation metal vapour<br />

generation at -196 °C was described as activation method for transition metal catalysts. 55 Iron<br />

vapour co-condensed with isoprene catalyzed the hydrosilylation with triethoxysilane. Not<br />

surprisingly exclusive 1,4-addition is observed.<br />

The group of Brunner investigated the influence of thermal or photoinduced activation of<br />

Fe(Cp)(CO) complexes in the hydrosilylation of acetophenone 4b with diphenylsilane<br />

forming quantitatively the silylated 1-phenylethanol 59 (Scheme 28). 56,57<br />

O<br />

4b<br />

+ Ph 2SiH 2<br />

0.3 mol% 65:<br />

neat, 23 h, hv<br />

0.5-1 mol% 62, 63, 64:<br />

neat, 50-80 °C, 24 h<br />

OSiHPh 2<br />

59<br />

OSiHPh 2<br />

60<br />

Cp`<br />

Cp`<br />

Fe<br />

Fe<br />

CO<br />

L<br />

CH3 61<br />

CO<br />

L<br />

C<br />

O<br />

62<br />

Cp`<br />

Fe Cp(OC)Fe<br />

CO<br />

OC<br />

CH3 63<br />

Scheme 28. Asymmetric hydrosilylation with chiral iron complexes.<br />

H<br />

Men 2<br />

P<br />

64<br />

CH 3<br />

Men = Menthyl<br />

Fe CO<br />

The absence of silyl enol ether 60 can be explained by a hydrosilylation mechanism involving<br />

a Fe-Si-H three-centre bond rather than a Fe-H species. Although optically active ligands of<br />

type 61, 62 and 63 containing chiral iron atoms were used, only poor enantioselectivities<br />

below 10% ee have been achieved (L = DIOP [O-isopropyliden-2,3-dihydroxy-1,4-<br />

57


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

bis(diphenylphosphino)butane]). 58 The rate determining step in this reaction is the thermal<br />

loss of the phosphine ligand L in 61 and 62 and methyl migration in 63, followed by addition<br />

of phenylsilane and subsequent reaction with acetophenone. Under photo-irradiation<br />

conditions hydrosilylation occured very fast with complex 64 as catalyst producing the<br />

silylated (S)-1-phenylethanol 59 in 33% ee. This was the first appreciable example for an<br />

enantioselective iron-catalyzed hydrosilylation of ketones. Very recently, Nishiyama et al.<br />

pursued a different elegant strategy to find an efficient catalytic system based on iron. 59 They<br />

combined Fe(OAc)2 and multi-nitrogen-based ligand such as N,N,N´,N´-<br />

tetramethylethylendiamine (tmeda), bis-tert-butyl-bipyridine (bipy-tb), or<br />

bis(oxazolinyl)pyridine (pybox) to catalyze the hydrosilylation of ketones to give the<br />

corresponding alcohol after acidic cleavage of the silyl ether (Scheme 29). The reaction was<br />

carried out under mild condition (THF at 65 °C, 24 hours) producing yields up to 95%. Using<br />

tmeda as nitrogen ligand hydrosilylation works also with other aromatic ketones.<br />

R<br />

OH<br />

5b: R = H yield: 93%<br />

5o: R = 4-MeO yield: 94%<br />

5t: R = 4-PhO yield: 90%<br />

OH<br />

5w: yield: 62%<br />

MeO<br />

OH OH<br />

OMe<br />

5p: yield: 94% 5u: yield: 72%<br />

OH<br />

5x: yield: 91%<br />

OH<br />

n-C 5H 11<br />

5y: yield: 94%<br />

5v: yield: 94%<br />

OH<br />

OH<br />

5k: yield: 89%<br />

Scheme 29. Substrate cope of the Fe-catalyzed hydrosilylation of ketones in the presence of<br />

tmeda.<br />

Ph<br />

5 mol% Fe(OAc) 2<br />

O nitrogene ligand<br />

OH<br />

(EtO) 2MeSiH<br />

*<br />

65<br />

THF, 65 °C, work<br />

up with H 3O +<br />

Ph<br />

66<br />

Bn<br />

O<br />

N<br />

N<br />

pybox-bn<br />

67<br />

N<br />

O<br />

Bn<br />

O<br />

N<br />

H<br />

N N<br />

R R<br />

58<br />

O<br />

68: R = i-Pr bopa-ip<br />

69: R = t-Bu bopa-tb<br />

Scheme 30. Asymmetric hydrosilylation of ketone 65 with iron catalysts according to<br />

Nishiyama and co-workers.


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

The application of chiral tridentate nitrogen ligands leads to the enantioselective reduction of<br />

methyl 4-phenylphenylketone 65. While pybox-bn 67 gives 37% ee of 66, bopa-ip 68 and tb<br />

69 increased the enantioselectivity up to 57% ee and 79% ee, respectively. Recent results<br />

from our group revealed that it is possible to perform Fe-catalyzed hydrosilylations of<br />

acetophenone also in the presence of chiral phosphine ligands. Here enantioselectivities up to<br />

80% ee have been achieved. 60<br />

Iron-catalyzed hydrosilylations of olefins in the presence of nitrogen ligands were first<br />

realized by Chirik et al. 34,35 They investigated the reaction of 1-hexene with PhSiH3 or Ph3SiH<br />

in the presence of iron catalyst 34 containing a tridentate pyridinediimine ligand (Scheme 17)<br />

which produce hydrosilylation over a course of minutes at ambient temperature. In both cases,<br />

the anti-Markovnikov product was formed exclusively. On the basis of these results a series of<br />

olefins (see Scheme 17) was examined. Terminal alkenes such as 1-hexene and styrene react<br />

most rapidly followed by gem-disubstituted olefins and internal olefins. The silylation of<br />

alkynes was also examined and proceeds efficiently under mild conditions to the<br />

corresponding silylalkene.<br />

3.6. Concluding Remarks<br />

Since the beginning of transition metal catalyzed hydrogenations and hydrosilylations more<br />

than 40 years ago a tremendous number of studies were directed to develop powerful methods<br />

for organic synthesis. In the past Rh, Ir and Ru clearly constitute the metals of choice for such<br />

transformations. Several hundreds of catalysts are nowadays commercially available for<br />

chemists around the world. However, the accelerated costs and the stronger requirements for<br />

pharmaceuticals cast a shadow on these commonly used transition metals.<br />

For the mid-term future we expect an increased use of more available and “biomimetic”<br />

metals. Clearly, iron is an ideal example for this. Unfortunately simple method transfer from<br />

Rh, Ir and Ru to Fe is not readily possible. So far iron based hydrogenations and<br />

hydrosilylations are far off to be general methods for organic chemistry. Typically<br />

comparably high catalyst loadings and harsh reaction conditions are required. In order to<br />

allow for more synthetic applications especially more functional group tolerance and also<br />

efficient control of stereoselectivity is needed. There are some promising developments like<br />

the recent work of Chirik and Nishiyama who demonstrates that these goals can be achieved.<br />

It will be interesting to take part in this renaissance of iron chemistry.<br />

59


3.7. References<br />

3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

1 a) P. T. Anastas, M. M. Kirchhoff, Acc. Chem. Res. 2002, 35, 686–694; b) P. T. Anastas, M.<br />

M. Kirchhoff, T. C. Williamson, Appl. Catal. A: Gen. 2001, 221, 3–13.<br />

2 a) H.-U. Blaser, B. Pugin, F. Spindler, J. Mol. Catal. 2005, 231, 1–20; b) H.-U. Blaser,<br />

Chem. Commun. 2003, 293–296; c) H.-U. Blaser in “Chiral Catalysis – Asymmetric<br />

Hydrogenations” supplement to chimica oggi 2004, 4–5.<br />

3 The price for iron is relatively low (current price for 1 t of iron ~300 US-dollar) in<br />

comparison to rhodium (5975 US-dollar per troy ounce), ruthenium (870 US-dollar per troy<br />

ounce) or iridium (460 US-dollar per troy ounce), due to the widespread abundance of iron<br />

in earth’s crust and the easy accessibility. Furthermore, iron is an essential trace element<br />

(daily dose for humans 5–28 mg) and plays an important role in an extensive number of<br />

biological processes.<br />

4 For an excellent review see: C. Bolm, J. Legros, J. Le Paih, L. Zani, Chem. Rev. 2004, 104,<br />

6217–6254.<br />

5 a) I. C. Lennon, J. A. Ramsden, Org. Process Res. Dev. 2005, 9, 110–112; b) J. M.<br />

Hawkins, T. J. N. Watson, Angew. Chem. 2004, 116, 3286–3290; c) R. Noyori, T. Ohkuma,<br />

Angew. Chem. 2001, 113, 40–75; d) M. Miyagi, J. Takehara, S. Collet, K. Okano, Org.<br />

Process Res. Dev. 2000, 4, 346–348; e) E. N. Jacobsen, A. Pfaltz, H. Yamamoto, (Eds.),<br />

Comprehensive Asymmetric Catalysis, Springer, Berlin, 1999; f) R. Noyori, Asymmetric<br />

Catalysis in Organic Synthesis, Wiley, New York, 1994.<br />

6 a) T. Ohkuma, D. Ishii, H. Takeno, R. Noyori, J. Am. Chem. Soc. 2007, in press; b) S.<br />

Sakuraba, K. Achiwa, Synlett 1991, 689–690; c) S. Sakuraba, N. Nakajima, K. Achiwa,<br />

Synlett 1992, 829–830; d) T. Kawaguchi, K. Saito, K. Matsuki, T. Iwakuma, M.Takeda,<br />

Chem. Pharm. Bull. 1993, 41, 639–642; e) T. Ohkuma, M. Koizuma, H. Doucet, T. Pham,<br />

M. Kozawa, K. Murata, E. Katayama, T. Yokozawa, T. Ikariya, R. Noyori, J. Am. Chem.<br />

Soc. 1998, 120, 13529–13530; f) T. Ohkuma, H. Ooka, R. Noyori, J. Am. Chem. Soc. 1995,<br />

117, 10417–10418; g) T. Ohkuma, H. Ikehira, T. Ikariya, R. Noyori, Synlett 1997, 467–468.<br />

7 a) Y. Takegami, Y. Watanabe, I. Kanaya, T. Mitsudo, T. Okajima, Y. Morishita, H. Masada,<br />

Bull. Chem. Soc. Jpn. 1968, 41, 2990–2994; b) R. Noyori, I. Umeda, T. Istigami, J. Org.<br />

Chem. 1972, 37, 1542–1545; c) Y. Watanabe, M. Yashita, T.-a. Mitsudo, M. Tanaka, Y.<br />

Takegami, Tetrahedron Lett. 1974, 22, 1879–1880; d) J. P. Collmann, R. G. Finke, P. L.<br />

Matlock, R. Wahren, J. I. Brauman, J. Am. Chem. Soc. 1976, 98, 4685–4687; e) M.<br />

Yamashita, K. Miyoshi, Y. Okada, R. Suemitsu, Bull. Chem. Soc. Jpn. 1982, 55, 1329–<br />

60


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

1330; f) more recently: Y. Moglie, F. Alonso, C. Vitale, M. Yus, G. Radivoy, Tetrahedron<br />

2006, 62, 2812–2819.<br />

8 L. Markó, M. A. Radhi, I. Ötvös, J. Organomet. Chem. 1981, 218, 369–376.<br />

9 L. Markó, J. Palágyi, Transition Met. Chem. 1983, 8, 207–209.<br />

10 a) K. Jothimony, S. Vancheesan, J. Mol. Catal. 1989, 52, 301–304; b) K. Jothimony, S.<br />

Vancheesan, J. C. Kuriacose, J. Mol. Catal. 1985, 32, 11–16.<br />

11 J.-S. Chen, L.-L. Chen, Y. Xing, G. Chen, W.-Y. Shen, Z.-R. Dong, Y.-Y. Li, J.-X. Gao,<br />

Huaxue Xuebao 2004, 62, 1745–1750.<br />

12 S. Enthaler, B. Hagemann, G. Erre, K. Junge, M. Beller, Chem. Asian. J. 2006, 1, 598–604.<br />

13 H. Zhang, C.-B. Yang, Y.-Y. Li, Z.-R. Donga, J.-X. Gao, H. Nakamura, K. Murata, T.<br />

Ikariya, Chem. Commun. 2003, 142–143.<br />

14 S. Enthaler, G. Erre, K. Junge, M. Beller, manuscript submitted.<br />

15 a) J. S. M. Samec, J.-E. Bäckvall, P. G. Andersson, P. Brandt, Chem. Soc. Rev. 2006, 35,<br />

237–248; b) T. Ikariya, K. Murata, R. Noyori, Org. Biomol. Chem. 2006, 4, 393–406 and<br />

references therein; c) H. Guan, M. Iimura, M. P. Magee, J. R. Norton, G. Zhu, J. Am. Chem.<br />

Soc. 2005, 127, 7805–7814; d) M. Gómez, S. Janset, G. Muller, G. Aullón, M. A. Maestro,<br />

Eur. J. Inorg. Chem. 2005, 4341–4351; e) K. Muñiz, Angew. Chem. 2005, 117, 6780–6785;<br />

Angew. Chem. Int. Ed. 2005, 44, 6622–6627; f) C. Hedberg, K. Källström, P. I. Arvidsson,<br />

P. Brandt, P. G. Andersson, J. Am. Chem. Soc. 2005, 127, 15083–15090; g) A. S. Y. Yim,<br />

M. Wills, Tetrahedron 2005, 61, 7994–8004; h) S. E. Clapham, A. Hadzovic, R. H. Morris,<br />

Coord. Chem. Rev. 2004, 248, 2201–2237; i) T. Koike, T. Ikariya, Adv. Synth. Catal. 2004,<br />

346, 37–41; j) P. Brandt, P. Roth, P. G. Andersson, J. Org. Chem. 2004, 69, 4885–4890; k)<br />

J.-W. Handgraaf, J. N. H. Reek, E. J. Meijer, Organometallics 2003, 22, 3150–3157; l) C. P.<br />

Casey, J. B. Johnson, J. Org. Chem. 2003, 68, 1998–2001; m) C. A. Sandoval, T. Ohkuma,<br />

K. Muñiz, R. Noyori, J. Am. Chem. Soc. 2003, 125, 13490–13503; n) R. Noyori, Angew.<br />

Chem. 2002, 114, 2108–2123; Angew. Chem. Int. Ed. 2002, 41, 2008–2022; o) C. P. Casey,<br />

S. W. Singer, D. R. Powell, R. K. Hayashi, M. Kavana, J. Am. Chem. Soc. 2001, 123, 1090–<br />

1100; p) O. Pàmies, J. E. Bäckvall, Chem. Eur. J. 2001, 7, 5052–5058; q) C. S. Yi, Z. He,<br />

Organometallics 2001, 20, 3641–3643; r) R. Noyori, M. Yamakawa, S. Hashiguchi, J. Org.<br />

Chem. 2001, 66, 7931–7944; s) M. Yamakawa, H. Ito, R. Noyori, J. Am. Chem. Soc. 2000,<br />

122, 1466–1478; t) D. A. Alonso, P. Brandt, S. J. M. Nordin, P. G. Andersson, J. Am.<br />

Chem. Soc. 1999, 121, 9580–9588; u) D. G. I. Petra, J. N. H. Reek, J.-W. Handgraaf, E. J.<br />

Meijer, P. Dierkes, P. C. J. Kamer, J. Brussee, H. E. Schoemaker, P. W. N. M. van<br />

Leeuwen, Chem. Eur. J. 2000, 6, 2818–2829; v) A. Aranyos, G. Csjernyik, K. J. Szabó, J.<br />

61


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

E. Bäckvall, Chem. Commun. 1999, 351–352; w) M. L. S. Almeida, M. Beller, G. Z. Wang,<br />

J. E. Bäckvall, Chem. Eur. J. 1996, 2, 1533–1536.<br />

16 S. Enthaler, G. Erre, M. K. Tse, K. Junge, M. Beller, Tetrahedron Lett. 2006, 47, 8095–<br />

8099.<br />

17 S. Enthaler, B. Spilker, G. Erre, M. K. Tse, K. Junge, M. Beller, manuscript submitted.<br />

18 a) W. S. Knowles, M. J. Sabacky, J. Chem. Soc. Chem. Commun. 1968, 1445–1446; b) W.<br />

S. Knowles, Angew. Chem. 2002, 114, 2097–2107; c) R. Selke, H. Pracejus, J. Mol. Catal.<br />

1986, 37, 213–225; d) R. Selke, J. Organomet. Chem. 1989, 370, 241–248.<br />

19 a) H. Pracejus, Koordinationschemische Katalyse organischer Reaktionen, Verlag Theodor<br />

Steinkopf, Dresden, 1977; b) Y. Takegami, T. Ueno, T. Fujii, Bull. Chem. Soc. Jpn. 1969,<br />

42, 1663–1667; c) J. W. Kalecic, F. K. Schmidt, Cinetica i Kataliz Moskva 1966, 7, 614–<br />

618; d) W. G. Lipovic, F. K. Schmidt, J. W. Kalecic, Cinetica i Kataliz Moskva 1967, 8,<br />

1300–1306.<br />

20 Y. Takegami, Y. Watanabe, I. Kanaya, T. Mitsudo, T. Okajima, Y. Morishita, H. Masada,<br />

Bull. Chem. Soc. Jpn. 1968, 41, 2990–2994 and references therein.<br />

21 E. N. Frankel, E. A. Emken, H. M. Peters, V. L. Davison, R. O. Butterfield, J. Org. Chem.<br />

1964, 29, 3292–3297.<br />

22 E. N. Frankel, E. A. Emken, V. L. Davison, J. Org. Chem. 1965, 30, 2739–2745.<br />

23 M. Cais, N. Maoz, J. Chem. Soc. (A) 1971, 1811–1820.<br />

24 Y. Tajima, E. Kunioka, J. Org. Chem. 1968, 33, 1689–1690.<br />

25 T. Nishiguchi, K. Fukuzumi, Bull. Chem. Soc. Jpn. 1972, 45, 1656–1660.<br />

26 M. A. Schroeder, M. S. Wrighton, J. Am. Chem. Soc. 1976, 98, 551–558.<br />

27 a) R. L. Whetten, K.-J. Fu, E. R. Grant, J. Am. Chem. Soc. 1982, 104, 4270–4272; b) M. E.<br />

Miller, E. R. Grant, J. Am. Chem. Soc. 1984, 106, 4635–4636; c) H. Nagorski, M. J.<br />

Mirbach, J. Organomet. Chem. 1985, 291, 199–204.<br />

28 H. Inoue, M. Sato, J. Chem. Soc. Chem. Commun. 1983, 983–984.<br />

29 H. Inoue, M. Suzuki, J. Chem. Soc. Chem. Commun. 1980, 817–818.<br />

30 Complex 26 furthermore exhibited activity in the hydrogenation of alkynes to yield alkenes.<br />

C. Bianchini, A. Mell, M. Peruzzini, P. Frediani, C. Bohanna, M. A. Esteruelas, L. A. Oro,<br />

Organometallics 1992, 11, 138–145.<br />

31 C. Bianchini, E. Farnetti, M. Graziani, M. Peruzzini, A. Polo, Organometallics 1993, 12,<br />

3753–3761.<br />

32 K. Kano, M. Takeuchi, S. Hashimoto, Z.-i. Yoshida, J. Chem. Soc. Chem. Commun. 1991,<br />

1728–1729.<br />

62


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

33 a) S. Sakaki, T. Sagara, T. Arai, T. Kojima, T. Ogata, K. Obkubo, J. Mol. Catal. 1992, 75,<br />

L33–L37; b) S. Sakaki, T.-i. Kojima, T. Arai, J. Chem. Soc. Dalton Trans. 1994, 7–11.<br />

34 S. C. Bart, E. Lobkovsky, P. J. Chirik, J. Am. Chem. Soc. 2004, 126, 13794–13807.<br />

35 A. M. Archer, M. W. Bouwkamp, M.-P. Cortez, E. Lobkovsky, P. J. Chirik,<br />

Organometallics 2006, 25, 4269–4278.<br />

36 R. J. Trovitch, E. Lobkovsky, P. J. Chirik, Inorg. Chem. 2006, 45, 7252–7260.<br />

37 S. C. Bart, E. J. Hawrelak, E. Lobkovsky, Organometallics 2005, 24, 5518–5527.<br />

38 M. A. Radhi, L. Markó, J. Organomet. Chem. 1984, 262, 359–364.<br />

39 T. J. Lynch, M. Banah, H. D. Kaesz, C. R. Porter, J. Org. Chem. 1984, 49, 1266–1270.<br />

40 K. Yamamoto, T. Hayashi in M. Beller, C. Bolm, (Eds.), Transition Metals for Organic<br />

Synthesis, Wiley-VCH, Weinheim, 2nd Ed. 2004.<br />

41 B. Marciniec, Comprehensive Handbook on Hydrosilylation, Pergamon Press, Oxford,<br />

1992.<br />

42 R. K. Freidlina, E. C. Chukovskaya, J. Tsao, A. N. Nesmeyanov, Dokl. Akad. Nauk. SSSR,<br />

1960, 132, 374–377.<br />

43 A. N. Nesmeyanov, R. Kh. Freidlina, E. C. Chukovskaya, R. G. Petrova, A. B. Belyavsky,<br />

Tetrahedron 1962, 17, 61–68.<br />

44 G. V. Nurtdinova, G. Gailiunas, V. P. Yur`ev, Izv. Akad. Nauk SSSR, Ser. Khim. 1981, 11,<br />

2652.<br />

45 G. Gailiunas, G. V. Nurtdinova, V. P. Yur`ev, G. A. Tolstikov, S. R. Rafikov, Izv. Akad.<br />

Nauk SSSR, Ser. Khim. 1982, 4, 914–920.<br />

46 I. V. Savos`kina, N. A. Kuz`min, G. G. Galust`yan, Izv. Akad. Nauk SSSR, Ser. Khim. 1990,<br />

2, 483–485.<br />

47 F. Kakiuchi, Y. Tanaka, N. Chatani, S. Murai, J. Orgmet. Chem. 1993, 456, 45–47.<br />

48 M. A. Schroeder, M. S. Wrighton, J. Organomet. Chem. 1977, 128, 345–358.<br />

49 I. Fischler, F.-W. Grevels, J. Organomet. Chem. 1980, 204, 181–190.<br />

50 C. U. Pittman Jr., W. D. Honnick, M. S. Wrighton, R. D. Sanner, R. G. Austin,<br />

Fundamental Research Homogeneous Catalysis 1979, 3, 603–619.<br />

51 R. G. Austin, R. S. Paonessa, P. J. Giordano, M. S. Wrighton, Adv. Chem. Ser. 1978, 168,<br />

189–214.<br />

52 C. L. Randolph, M. S. Wrighton, J. Am. Chem. Soc. 1986, 108, 3366-3374.<br />

53 G. Bellachioma, G. Cardaci, E. Colomer, R. J. P. Corriu, A. Vioux, Inorg. Chem. 1989, 28,<br />

519–525.<br />

63


3. Reduction of unsaturated compounds with homogeneous iron catalysts<br />

54 C. U. Pittman Jr., M. G. Richmond, M. Absi-Halabi, H. Beurich, F. Richter, H.<br />

Vahrenkamp, Angew. Chem. 1982, 94, 805-806; Angew. Chem. Int. Ed. Engl. 1982, 21,<br />

786-787.<br />

55 A. J. Cornish, M. F. Lappert, J. J. MacQuitty, R. K. Maskell, J. Organomet. Chem. 1979,<br />

177, 153–161.<br />

56 H. Brunner, K. Fisch, Angew. Chem. 1990, 102, 1189–1191; Angew. Chem. Int. Ed. Engl.<br />

1990, 29, 1131–1132. Brunner et al. reported the successful application of<br />

[(indenyl)Fe(CO)2(CH3)] complexes in hydrosilylation of ketones, but extension to<br />

hydrogenation of styrenes with molecular hydrogen was impossible since catalyst<br />

deactivation was observed.<br />

57 H. Brunner, K. Fisch, J. Organomet. Chem. 1991, 412, C11–C13.<br />

58 H. Brunner, R. Eder, B. Hammer, U. Klement, J. Organomet. Chem. 1990, 394, 555–567.<br />

59 H. Nishiyama, A. Furuta, Chem. Commun. 2007, 7, 760–762.<br />

60 N. Shaik, S. Enthaler, K. Junge, M. Beller, unpublished results.<br />

64


4. Objectives of the work 65<br />

4. Objectives of the work<br />

4.1. Application of monodentate phosphine ligands in asymmetric hydrogenations<br />

Since earlier works have proven an excellent behaviour of ligands containing 4,5–dihydro–<br />

3H–dinaphtho[2,1–c;1´,2´–e]phosphepine scaffold in the asymmetric hydrogenation of α–<br />

amino acid precursors, itaconic acid derivatives and β–ketoesters, we focused our attention on<br />

extension of scope and limitation of this ligand tool box. Several challenging substrate<br />

classes, based on unsaturated C-C bonds, were synthesized and subjected to catalytic<br />

reduction.<br />

4.1.1. Asymmetric hydrogenation of N–acyl enamides<br />

Until the end of 2004 only two reports were published dealing with the rhodium–catalyzed<br />

hydrogenation of N–acyl enamides in the presence of monodentate phosphines. On the one<br />

hand the utilization of chiral phospholanes by Fiaud and co–workers, which attained good<br />

enantioselectivities up to 73% ee, when (Z)–N–(1–phenylprop–1–enyl)acetamide was<br />

hydrogenated. 1 On the other hand, Reetz et al. described the rhodium–catalyzed<br />

hydrogenation of N–(1–phenylvinyl)–acetamide with ligands based on 4,5–dihydro–3H–<br />

dinaphtho[2,1–c;1´,2´–e]phosphepine (vide supra). 2 Surprisingly, the tert–butyl–substituted<br />

ligand showed better selectivity and yield than the phenyl–substituted ligand, which is in<br />

contrast to other benchmark tests such as the asymmetric hydrogenation of α–amino acid<br />

precursors, dimethyl itaconate and β–ketoesters where always the opposite behaviour was<br />

observed. The interesting results of Reetz and co–workers stimulated us to study the potential<br />

of our ligand library in the field of asymmetric hydrogenation of N–acyl enamides in more<br />

detail.<br />

4.1.2. Asymmetric hydrogenation of β–dehydroamino acid derivatives<br />

Since, β–amino acids are interesting building blocks for the synthesis of biologically active<br />

compounds the synthesis via asymmetric hydrogenation of unsaturated precursors offers a<br />

versatile and powerful approach. Currently bidentate phosphorous ligands have been proven<br />

to be effective in the asymmetric hydrogenation of β–dehydroamino acid derivatives with<br />

enantioselectivities up >99% ee, while a monodentate version is so far scarcely reported. Here<br />

only the group of Börner accounted enantioselectivities up to 93% ee with monodentate


4. Objectives of the work 66<br />

phospholane ligand. 3 We became interested in studying the behaviour of chiral monodentate<br />

44,5–dihydro–3H–dinaphtho[2,1–c;1´,2´–e]phosphepine in the rhodium–catalyzed<br />

asymmetric hydrogenation of various β–dehydroamino acid derivatives to give optically<br />

active β–amino acids.<br />

4.1.3. Asymmetric hydrogenation of enol carbamates<br />

Pioneering work in the field of enantioselective hydrogenation of enol carbamates, which are<br />

suitable precursors for chiral alcohols, has been reported by Feringa, de Vries, Minnaard and<br />

co–workers applying rhodium–catalysts containing monodentate phosphoramidites<br />

(MonoPhos–family) and obtaining enantioselectivities up to 98% ee. 4 However, until today<br />

no additional studies on the reduction of enol carbamates were reported. Hence, we became<br />

interested in testing our ligand library in this purpose.<br />

4.2. Transfer hydrogenation<br />

4.2.1. Homogeneous iron catalyst in reduction chemistry<br />

With regard to future catalyst developments a fundamental challenge will be the<br />

substitution of expensive and rare transition metals by inexpensive and abundantly available<br />

metals such as iron. So far, homogeneous iron catalysts have been most frequently applied for<br />

carbon–carbon coupling reactions, such as olefin polymerizations, cross couplings, and<br />

cycloadditions. However, much less attention was directed towards iron–catalyzed (transfer)<br />

hydrogenations, albeit the relevance of such reductions is evident even with respect to<br />

industrial applications. Until today only a few number of research groups reported the<br />

application of iron salts and iron complexes in the reduction of α,β–unsaturated carbonyl<br />

compounds and ketones using a transfer hydrogenation protocol. 5 For tuning the activity and<br />

selectivity of these catalysts oxygen and moisture sensitive tetradentate phosphines or<br />

aminophosphines have been predominantly applied as ligands. Our goal was to develop a<br />

practical iron hydrogenation catalyst system, which should be easily prepared and highly<br />

tuneable. Thus, the use of commercially available iron complexes in combination with two<br />

different ligands (phosphines and amines) instead of tetradentate ligands seemed to be a<br />

useful approach. Furthermore inspired by nature we thought that multidentate nitrogen ligands<br />

should be also suitable ligands for stabilizing iron as metal center in transfer hydrogenations.<br />

The influence of different reaction parameters on the catalytic activity will be investigated in


4. Objectives of the work 67<br />

the transfer hydrogenation of aliphatic and aromatic ketones utilizing 2–propanol as hydrogen<br />

donor in the presence of base.<br />

4.2.2. Transfer hydrogenation of ketones with ruthenium catalysts<br />

4.2.2.1. Transfer hydrogenation with ruthenium carbene catalysts<br />

With regard to transfer hydrogenations different carbene or carbene-phosphine–systems<br />

containing Rh, Ir, Ru or Ni have been reported. Excellent turnover frequencies up to 120000<br />

h -1 were reported by the groups of Baratta and Herrmann applying a ruthenium–carbene–<br />

phosphine–catalyst. 6 However, for reduction of a typical substrate, e.g. acetophenone, with<br />

phosphine–free ruthenium–carbene catalysts lower turnover frequencies (TOF 333 h -1 ) 7 were<br />

achieved in comparison to iridium (500 h -1 ) 8 and rhodium systems (583 h -1 ). 9 Due to the<br />

economical benefit of ruthenium metal compared to rhodium or iridium and the advantages of<br />

phosphine–free systems, it will be an important goal to search for more active ruthenium<br />

carbene catalysts.<br />

4.2.2.2. Asymmetric transfer hydrogenation with ruthenium catalysts<br />

Significantly less is known of the transfer of chiral information for tridentate ligands, while<br />

transfer hydrogenation is dominated by highly active bidentate systems. 10 Up to now only a<br />

limited number of auspicious tridentate nitrogen-containing N,N,N–ligands were established.<br />

For example (R)–phenyl–ambox and different pyridinebisoxazoline (pybox) ligands have<br />

been applied for the reduction of acetophenone. More recently, a new class of chiral tridentate<br />

amines so called pybim (parent structure: 2,6-bis–([4R,5R]–4,5–diphenyl–4,5–dihydro–1H–<br />

imidazol–2-yl)–pyridine) for the ruthenium–catalyzed epoxidations was introduced by Beller<br />

et al. 11 The resemblance between pybim and pybox stimulated our research to study the<br />

potential of this class of ligands in the transfer hydrogenation of aromatic and aliphatic<br />

ketones.


4. Objectives of the work 68<br />

4.3. References<br />

1 C. Dobrota, M. Toffano, J.-C. Fiaud, Tetrahedron Lett. 2004, 45, 8153–8156.<br />

2 M. T. Reetz, G. Mehler, A. Meiswinkel, Tetrahedron: Asymmetry 2004, 15, 2165–2167.<br />

3 V. Bilenko, A. Spannenberg, W. Baumann, I. Komarov, A. Börner, Tetrahedron:<br />

Asymmetry 2006, 17, 2082–2087.<br />

4 a) X. Jiang, M. van den Berg, A. J. Minnaard, B. Feringa, J. G. de Vries, Tetrahedron:<br />

Asymmetry 2004, 15, 2223–2229; b) L. Panella, B. L. Feringa, J. G. de Vries, A. J.<br />

Minnaard, Org. Lett. 2005, 7, 4177–4180.<br />

5 a) K. Jothimony, S. Vancheesan, J. Mol. Catal. 1989, 52, 301–304; b) K. Jothimony, S.<br />

Vancheesan, J. C. Kuriacose, J. Mol. Catal. 1985, 32, 11–16; c) J.-S. Chen, L.-L. Chen,<br />

Y. Xing, G. Chen, W.-Y. Shen, Z.-R. Dong, Y.-Y. Li, J.-X. Gao, Huaxue Xuebao 2004,<br />

62, 1745–1750; d) C. Bianchini, E. Farnetti, M. Graziani, M. Peruzzini, A. Polo,<br />

Organometallics 1993, 12, 3753–3761.<br />

6 W. Baratta, J. Schütz, E. Herdtweck, W. A. Herrmann, P. Rigo, J. Organomet. Chem.<br />

2005, 690, 5570–5575.<br />

7 A. A. Danopoulos, S. Winston, W. B. Motherwell, Chem. Commun. 2002, 1376–1377.<br />

8 M. Albrecht, J. R. Miecznikowski, A. Samuel, J. W. Faller, R. H. Crabtree,<br />

Organometallics 2002, 21, 3596–3604.<br />

9 M. Poyatos, E. Mas-Marzá, J. A. Mata, M. Sanaú, E. Peris, Eur. J. Inorg. Chem. 2003,<br />

1215–1221.<br />

10 a) Y. Jiang, Q. Jiang, X. Zhang, J. Am. Chem. Soc. 1998, 120, 3817–3818; b) D.<br />

Cuervo, M. P. Gamasa, J. Gimeno, Chem. Eur. J. 2004, 10, 425–432.<br />

11 a) S. Bhor, G. Anilkumar, M. K. Tse, M. Klawonn, C. Döbler, B. Bitterlich, A.<br />

Grotevendt, M. Beller, Org. Lett. 2005, 7, 3393–3396; b) G. Anilkumar, S. Bhor, M. K.<br />

Tse, M. Klawonn, B. Bitterlich, M. Beller, Tetrahedron: Asymmetry 2005, 16, 3536–<br />

3561.


5. Application of monodentate phosphine ligands in asymmetric hydrogenation 69<br />

5.1. Enantioselective rhodium–catalyzed hydrogenation of enamides in<br />

the presence of chiral monodentate phosphines<br />

Stephan Enthaler, Bernhard Hagemann, Kathrin Junge, Giulia Erre, and Matthias<br />

Beller*, Eur. J. Org. Chem. 2006, 2912–2917.<br />

Contributions<br />

Ligands 5c and 5f were synthesized by Dr. B. Hagemann. Furthermore, the asymmetric<br />

hydrogenation of substrate 6f and the combinatorial ligand approach (Table 3) were carried by<br />

Dr. B. Hagemann.


FULL PAPER<br />

DOI: 10.1002/ejoc.200600024<br />

Enantioselective Rhodium-Catalyzed Hydrogenation of Enamides in the<br />

Presence of Chiral Monodentate Phosphanes<br />

Stephan Enthaler, [a] Bernhard Hagemann, [a] Kathrin Junge, [a] Giulia Erre, [a] and<br />

Matthias Beller* [a]<br />

Keywords: Asymmetric catalysis / Homogeneous catalysis / Hydrogenation / Monodentate phosphane ligands / Enamides<br />

The rhodium-catalyzed asymmetric hydrogenation of different<br />

acyclic and cyclic N-acyl enamides to give N-acyl-protected<br />

optically active amines has been examined for the first<br />

time in detail in the presence of chiral monodentate 4,5-dihydro-3H-dinaphthophosphepines<br />

5a–i. The enantioselectivity<br />

is largely dependent on the nature of the substituent at the<br />

Introduction<br />

The importance of chiral amines is demonstrated by their<br />

broad scope of applications as building blocks and synthons<br />

for pharmaceuticals and agrochemicals. In addition,<br />

they are of significant use for the synthesis of natural compounds,<br />

chiral auxiliaries, and optically active ligands. [1]<br />

Hence, the development of new and improved methods for<br />

the preparation of enantiomerically pure amines is an important<br />

target of industrial and academic research. Several<br />

synthetic approaches to optically active amines have been<br />

established in the past. Still one of the most practical approaches<br />

is the separation of enantiomers through resolution,<br />

e.g. in the presence of chiral acids. The main drawback<br />

of this concept is the poor atom efficiency and the negative<br />

environmental impact. A more attractive access is represented<br />

by applying enantiomerically pure starting materials<br />

from the “chiral pool”. Another powerful strategy is the<br />

application of stereoselective syntheses in particular reactions<br />

catalyzed by transition-metal complexes. [2] Examples<br />

of such transition-metal-catalyzed reactions, which offer a<br />

versatile and elegant approach to enantiomerically pure<br />

amines, are the alkylation or arylation of imines, hydrocyanation<br />

of carbon–carbon double bonds and aziridination.<br />

[2c] However, the most popular catalytic reaction with<br />

respect to industrial applications is the asymmetric hydrogenation<br />

of imines or N-protected enamines and enamides.<br />

[3]<br />

Pioneering work in the field of enantioselective hydrogenation<br />

of enamides has been reported by Kagan and co-<br />

[a] Leibniz-Institut für Katalyse an der <strong>Universität</strong> <strong>Rostock</strong> e.V.,<br />

Albert-Einstein-Str. 29A, 18059 <strong>Rostock</strong>, Germany<br />

Fax: +49-381-12815000<br />

E-mail: matthias.beller@ifok-rostock.de<br />

2912<br />

phosphorus atom and the enamide substrate. Applying optimized<br />

conditions up to 95% ee and catalyst activity up to<br />

2000 h –1 (TOF) have been achieved.<br />

(© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim,<br />

Germany, 2006)<br />

workers. Applying diphosphane rhodium catalysts enantioselectivity<br />

up to 90% has been obtained. [4] Over a period of<br />

nearly 30 years the development of new ligands for asymmetric<br />

hydrogenation has focused on the synthesis of bidentate<br />

ligands. During this time several effective bidentate<br />

phosphanes such as Me-DuPhos, Me-BPE, DIOP, BINAP<br />

and others were developed for the hydrogenation of enamides.<br />

[4,5] At the end of the 20 th century the predominant<br />

role of bidentate ligands changed and monodentate ligands<br />

received more attention. [6,7] The advantages of monodentate<br />

phosphorus ligands are their easier synthesis and<br />

tunability compared to bidentate phosphanes. Scheme 1 illustrates<br />

some examples of monodentate phosphorus ligands<br />

for asymmetric hydrogenation on the basis of the chiral<br />

1,1�-binaphthol skeleton. Important contributions in<br />

this area came independently from Feringa and de Vries et<br />

al. [8] (phosphoramidites 1), Reetz et al. [9] (phosphites 2),<br />

and Pringle and Claver et al. [10] (phosphonites 3). Furthermore,<br />

excellent enantioselectivity was shown by Zhou and<br />

co-workers applying monodentate spiro phosphoramidites<br />

(SIPHOS). [11]<br />

Following the original work of Gladiali [12] and parallel<br />

to Zhang, [13] we have established the synthesis of various<br />

monodentate phosphanes based on a 4,5-dihydro-3H-dinaphtho[2,1-c;1�,2�-e]phosphepine<br />

framework (4 and 5a–i),<br />

which resembles the 1,1�-binaphthol core. [14] Recently, we<br />

have shown that asymmetric hydrogenation of amino acid<br />

precursors, dimethyl itaconate and β-keto esters proceeds<br />

with enantioselectivities up to 95% ee in the presence of<br />

ligand 5a. [13] Moreover, other groups demonstrated the usefulness<br />

of these ligands in several catalytic asymmetric reactions.<br />

[15]<br />

To the best of our knowledge there exists only one publication<br />

dealing with the hydrogenation of enamides in the<br />

© 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Eur. J. Org. Chem. 2006, 2912–2917


Enantioselective Hydrogenation of Enamides in the Presence of Phosphanes FULL PAPER<br />

Scheme 1. Chiral monodentate ligands with 2,2�-binaphthyl framework.<br />

presence of monodentate phosphepines. Here, Reetz et<br />

al. [16] described the rhodium-catalyzed hydrogenation of N-<br />

(1-phenylvinyl)acetamide (6a) with ligands 5a [conversion:<br />

50%; enantiomeric excess: 14% (R)] and 5i [conversion:<br />

94%; enantiomeric excess: 24% (R)] in dichloromethane. To<br />

our surprise the tert-butyl-substituted ligand 5i showed better<br />

selectivity and yield than the phenyl-substituted ligand<br />

5a. In our benchmark tests, for instance, in the asymmetric<br />

hydrogenation of amino acid precursors, dimethyl itaconate<br />

and β-keto esters, we always observed the opposite behaviour.<br />

Discussion<br />

The interesting results of Reetz and co-workers [15] stimulated<br />

us to study the potential of our ligand library 5a–i in<br />

the field of asymmetric hydrogenation of enamides in more<br />

detail. In general, the ligands are prepared by metallation of<br />

2,2�-dimethylbinaphthyl with two equiv. of n-butyllithium,<br />

followed by quenching with diethylamino(dichloro)phosphane.<br />

Deprotection with gaseous HCl produced 4-chloro-<br />

4,5-dihydro-3H-dinaphtho[2,1-c;1�,2�-e]phosphepine in a<br />

yield of 80%. This enantiomerically pure chlorophosphane<br />

is easily coupled with various Grignard reagents to give a<br />

broad selection of ligands of type 5.<br />

Unsubstituted and substituted N-(1-phenylvinyl)acetamides<br />

6a–6e were obtained by reacting the corresponding<br />

benzonitrile with a methyl Grignard followed by addition<br />

of acetic anhydride. [8e] The cyclic N-acyl enamide 6f was<br />

synthesized by a standard protocol implying transformation<br />

of the ketone to the corresponding oxime and subsequent<br />

reductive acylation with iron in the presence of acetic acid<br />

anhydride. [17]<br />

Initial studies on the influence of reaction conditions<br />

were carried out with N-(1-phenylvinyl)acetamide (6a) as<br />

substrate and 4-phenyl-4,5-dihydro-3H-dinaphtho[2,1c;1�,2�-e]phosphepine<br />

(5a) as our standard ligand. Typically,<br />

we used an in situ pre-catalytic mixture of 1 mol-%<br />

[Rh(cod) 2]BF 4 and 2.1 mol-% of the corresponding ligand.<br />

All hydrogenation reactions were carried out in an 8fold<br />

parallel reactor array with 3.0 mL reactor volume. [18]<br />

At first, we focused our attention on the influence of<br />

different solvents, such as dichloromethane, methanol, ethyl<br />

acetate, toluene, and also variation of the initial pressure<br />

(1.0 bar, 5.0 bar and 10 bar). Selected results are presented<br />

in Figure 1.<br />

Figure 1. Solvent and pressure variation. Reactions were carried<br />

outat30°C for 24 h with 0.0024 mmol [Rh(cod) 2]BF 4, 0.005 mmol<br />

ligand 5a and 0.24 mmol substrate in 2.0 mL solvent. Conversion<br />

and ee values were determined by GC [50 m Lipodex E (Macherey–<br />

Nagel), 80 °C, (R)-7a 26.7 min, (S)-7a 28.3 min]. The absolute configuration<br />

was determined by comparing the sign of specific rotation<br />

with reported data and the original sample (Aldrich). [5i]<br />

For all solvents best enantioselectivity (80–90% ee) is<br />

achieved at 1.0 bar, but without complete conversion within<br />

reasonable time (24 h). Increasing the pressure of hydrogen<br />

to 5.0 bar or 10.0 bar accelerated the reaction and <strong>full</strong> conversion<br />

is reached, but at somewhat lower selectivity. The<br />

results also indicated toluene as the solvent of choice for<br />

our model reaction (conversion: 96%; enantioselectivity:<br />

91%).<br />

To find the optimum of the enantioselectivity−pressure<br />

dependency we performed a fine tuning of the hydrogen<br />

pressure in toluene. At 2.5 bar hydrogen pressure both good<br />

enantioselectivity (93% ee) and acceptable reaction time<br />

(6 h) have been achieved.<br />

Next, we investigated the influence of the temperature on<br />

the selectivity and yield as described in Figure 2. Applying<br />

our model ligand 5a there is no pronounced effect on the<br />

selectivity between 10 and 30 °C. [19] At higher temperatures<br />

a decrease of the ee is observed (87% ee at 50 °C and 83%<br />

ee at 70 °C).<br />

Eur. J. Org. Chem. 2006, 2912–2917 © 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.eurjoc.org 2913


FULL PAPER<br />

Figure 2. Dependency of enantioselectivity vs. temperature. Reactions<br />

were carried out at the corresponding temperature for 1–24 h<br />

with 0.0024 mmol [Rh(cod) 2]BF 4, 0.005 mmol ligand 5a and<br />

0.24 mmol substrate in 2.0 mL toluene. Conversion and ee values<br />

were determined by GC [50 m Lipodex E (Macherey–Nagel),<br />

80 °C, (R)-7a 26.7 min, (S)-7a 28.3 min]. [5i]<br />

To estimate the activity of the catalyst in more detail the<br />

ratio between metal and substrate was varied starting from<br />

initially 1:100 to 1:2000. At 50 °C and 2.5 bar of hydrogen<br />

complete conversion was observed within one hour, which<br />

corresponds to a turnover frequency (TOF) of 2000 h –1 .<br />

In order to evaluate the substituent effect at the phosphorus<br />

atom of the ligand, we studied the asymmetric hydrogenation<br />

of N-(1-phenylvinyl)acetamide (6a) in the presence<br />

of nine different 4,5-dihydro-3H-dinaphtho[2,1-c;1�,2�e]phosphepines<br />

(5a–5i) using the optimized reaction conditions<br />

(2.5 bar hydrogen pressure, toluene, 30 °C, 6 h). These<br />

results are presented in Table 1.<br />

In agreement with our previous studies, it appeared<br />

that for obtaining good enantioselectivity aryl-substituted<br />

phosphepine ligands are necessary, while alkyl-substituted<br />

ligands showed significantly lower selectivity. Substitution<br />

on the phenyl ring at the phosphorus atom with electronwithdrawing<br />

(5b, 5c) as well as electron-donating groups<br />

(5d–5g) resulted in decreased selectivity in comparison to<br />

the phenyl ligand 5a. Here, only ligand 5e represented an<br />

exception (Table 1, Entry 5), which led to 93% ee.<br />

Interestingly, the behaviour of the tert-butyl ligand 5i is<br />

quite different in the hydrogenation of 6a. In contrast to<br />

the results obtained by Reetz et al., under our reaction conditions<br />

compound 5i induced a change in the absolute configuration<br />

of the product to the opposite (S)-enantiomer,<br />

albeit with significantly lower ee.<br />

In addition to the ligands 5, we also tested the commercially<br />

available (S)-MonoPhos ligand (1, R 1 =R 2 = Me),<br />

which gives excellent enantioselectivity in various hydrogenations<br />

of prochiral double bonds. [8,20] However, in the present<br />

test reaction we obtained only poor yield and enantioselectivity<br />

(Table 1, Entry 10).<br />

2914<br />

S. Enthaler, B. Hagemann, K. Junge, G. Erre, M. Beller<br />

Table 1. Hydrogenation of N-(1-phenylvinyl)acetamide (6a). [a]<br />

Entry Ligand Conversion [%] [b] ee [%] [b,c]<br />

1 5a �99 93 (R)<br />

2 5b �99 72 (R)<br />

3 5c �99 88 (R)<br />

4 5d �99 87 (R)<br />

5 5e �99 93 (R)<br />

6 5f �99 63 (R)<br />

7 5g �99 90 (R)<br />

8 5h �99 30 (R)<br />

9 5i �99 16 (S)<br />

10 1 4 19 (S)<br />

[a] All reactions were carried out at 30 °C under 2.5 bar pressure<br />

of hydrogen for 6 h with 0.0024 mmol [Rh(cod) 2]BF 4, 0.005 mmol<br />

ligand and 0.24 mmol substrate in toluene (2.0 mL). [b] Conversion<br />

and ee values were determined by GC [50 m Lipodex E (Macherey–<br />

Nagel), 80 °C, (R)-7a 26.7 min, (S)-7a 28.3 min]. [c] The absolute<br />

configuration was determined by comparing the sign of specific<br />

rotation with reported data and the original sample (Aldrich). [5i]<br />

To demonstrate the scope and limitations of our ligand<br />

toolbox we performed the asymmetric hydrogenation of<br />

four different N-(1-phenylvinyl)acetamides (Table 2). Regarding<br />

the substrates both electron-donating substituents<br />

at the phenyl group, in particularly a methoxy group in<br />

para- as well as in meta-position and electron-withdrawing<br />

substituents such as para-fluoro or para-trifluoromethyl<br />

have been tested. Enantioselectivity up to 95% ee is obtained<br />

with ligands 5a (Table 2, Entry 1) and 5e (Table 2,<br />

Entry 5) for electron-rich substrates. In general, aryl phosphepines<br />

with electron-donating substituents showed better<br />

selectivity than phosphepines having electron-withdrawing<br />

groups.<br />

The introduction of electron-withdrawing substituents in<br />

N-(1-phenylvinyl)acetamide 6d and 6e decreased the<br />

enantioselectivity compared to the non-substituted N-(1phenylvinyl)acetamide<br />

(6a). Here, the best selectivities of<br />

83–86% ee are obtained with ligands 5a, 5e and 5g.<br />

Again for all substrates 6a–6e the alkyl-substituted 4,5dihydro-3H-dinaphtho[2,1-c;1�,2�-e]phosphepines<br />

gave significantly<br />

lower selectivity.<br />

Next, the hydrogenation of the sterically more hindered<br />

N-(3,4-dihydro-1-naphthyl)acetamide (6f) was tested<br />

(Scheme 2). The use of different pre-catalysts including<br />

[Rh(5a) 2(cod)]BF 4, [Rh(5e) 2(cod)]BF 4 and [Rh(5i) 2(cod)]<br />

BF 4 gave an excellent yield (�99%) of 7f.<br />

However, the observed enantioselectivity is significantly<br />

lower [5a: 15%(S); 5e: 11%(S); 5i: 35%(R)] when compared<br />

to 1,1-disubstituted enamides. These results indicate<br />

a crucial influence of the substitution pattern on the enamide<br />

group.<br />

It is important to note that in comparison to bidentate<br />

ligands the application of monodentate ligands offers an<br />

www.eurjoc.org © 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Eur. J. Org. Chem. 2006, 2912–2917


Enantioselective Hydrogenation of Enamides in the Presence of Phosphanes FULL PAPER<br />

Table 2. Asymmetric hydrogenation of substituted N-(1-phenylvinyl)acetamides<br />

6b–6e. [a]<br />

7b 7c<br />

Entry Ligand (R 1 =CH 3O; R 2 =H) (R 1 =H;R 2 =CH 3O)<br />

ee [%] [b,c] ee [%] [b,d]<br />

1 5a 91 (R) 95 (R)<br />

2 5b 63 (R) 72 (R)<br />

3 5c 79 (R) 84 (R)<br />

4 5d 87 (R) 87 (R)<br />

5 5e 92 (R) 89 (R)<br />

6 5f 33 (R) 53 (R)<br />

7 5g 89 (R) 89 (R)<br />

8 5h rac 26 (R)<br />

9 5i 21 (S) 23 (S)<br />

Entry Ligand 7d (R 1 =F) 7e (R 1 =CF 3)<br />

ee [%] [b,c] ee [%] [b,c]<br />

10 5a 86 (R) 78 (R)<br />

11 5b 51 (R) 62 (R)<br />

12 5c 71 (R) 73 (R)<br />

13 5d 71 (R) 65 (R)<br />

14 5e 86 (R) 78 (R)<br />

15 5f 34 (R) 68 (R)<br />

16 5g 83 (R) 85 (R)<br />

17 5h 4(R) 44 (R)<br />

18 5i rac 25 (S)<br />

[a] All reactions were carried out at 30 °C under 2.5 bar pressure<br />

of hydrogen for 6 h with 0.0024 mmol [Rh(cod) 2]BF 4, 0.005 mmol<br />

ligand and 0.24 mmol substrate in 2.0 mL toluene. [b] Conversion<br />

(�99 %) was determined by GC (Agilent Technologies, 30 m, 50–<br />

300 °C), ee values were determined by GC [50 m Lipodex E (Macherey–Nagel),<br />

100–180 °C, (R)-7b 36.1 min, (S)-7b 36.4 min, (R)-7c<br />

34.8 min, (S)-7c 35.0 min, (R)-7d 31.0 min, (S)-7d 31.2 min, (R)-<br />

7e 32.4 min, (S)-7e 32.7 min]. [c] The absolute configurations were<br />

determined by comparing the sign of specific rotation with reported<br />

data. [5i] [d] Absolute configuration of product was assigned<br />

by analogy.<br />

Scheme 2. Asymmetric hydrogenation of N-(3,4-dihydro-1-naphthyl)acetamide<br />

(6f). All reactions were carried out at 30 °C under<br />

2.5 bar pressure of hydrogen for 6 h with 0.0024 mmol [Rh(cod) 2]-<br />

BF 4, 0.005 mmol ligand and 0.24 mmol substrate in 2.0 mL toluene.<br />

Conversion was determined by GC (Agilent Technologies,<br />

30 m, 50–300 °C) and ee values were determined by HPLC [AD-<br />

H Chiralpak (Agilent Technologies), n-hexane/ethanol, 95:5, rate<br />

1.0 mL/min, (S)-7f 6.2 min, (R)-7f 7.2 min]. The absolute configuration<br />

was determined by comparing the sign of specific rotation<br />

with reported data.<br />

opportunity to replace one equiv. of the ligand by other chiral<br />

or achiral monodentate ligands. This remarkable combinatorial<br />

approach was introduced by Reetz et al., [9d–9f,15,21]<br />

and Feringa et al. [22] for different transition-metal-catalyzed<br />

reactions. [23]<br />

We followed this concept and used 4-phenyl-4,5-dihydro-<br />

3H-dinaphtho[2,1-c;1�,2�-e]phosphepine (5a) with different<br />

achiral phosphanes in a ratio of 1.05 to 1.05 equiv. in the<br />

presence of 1.0 equiv. of [Rh(cod) 2]BF 4. The results obtained<br />

for the hydrogenation of N-(1-phenylvinyl)acetamide<br />

(6a) under optimized reaction conditions are presented in<br />

Table 3.<br />

Table 3. Application of ligand mixtures in the asymmetric hydrogenation<br />

of N-(1-phenylvinyl)acetamide (6a). [a]<br />

Entry Ligand B Conversion [%] [b] ee [%] [b,c]<br />

1 5a �99 93 (R)<br />

2 PPh 3 �99 85 (R)<br />

3 P(p-MeO-C 6H 4) 3 �99 88 (R)<br />

4 P(p-Me-C 6H 4) 3 �99 81 (R)<br />

5 PCy 3 �99 74 (R)<br />

[a] All reactions were carried out at 30 °C under 2.5 bar pressure<br />

of hydrogen for 6 h with 0.0024 mmol [Rh(cod) 2]BF 4, 0.0025 mmol<br />

ligand 5a, 0.0025 mmol achiral ligand and 0.24 mmol substrate in<br />

2.0 mL toluene. [b] Conversion and ee values determined by GC<br />

[50 m Lipodex E (Macherey–Nagel), 80 °C, (R)-7a 26.7 min, (S)-7a<br />

28.3 min]. [c] The absolute configuration was determined by comparing<br />

the sign of specific rotation with reported data and the original<br />

sample (Aldrich). [5i]<br />

In general, we observed a slight decrease in enantioselectivity<br />

from that observed when using 2.1 equiv. of ligand<br />

5a. On the assumption that probably three different<br />

metal complexes could be formed, in particular, two homocombination<br />

products, the chiral [Rh(5a)(5a)(cod)]BF 4 and<br />

the achiral [Rh(L)(L)(cod)]BF 4, and the hetero-combination<br />

product [Rh(L)(5a)(cod)]BF 4, we deduce a significantly<br />

lower reaction rate or inactivity for the achiral complex<br />

[Rh(L)(L)(cod)]BF 4 relative to complex [Rh(5a)(5a)(cod)]-<br />

BF 4 or [Rh(L)(5a)(cod)]BF 4. Tricyclohexylphosphane<br />

(Table 3, Entry 5) was found to be the only exception, as<br />

we noticed a more pronounced decrease of the enantiomeric<br />

excess.<br />

Conclusions<br />

We demonstrated the application of monodentate 4,5-dihydro-3H-dinaphtho[2,1-c;1�,2�-e]phosphepines<br />

5 in the<br />

rhodium-catalyzed asymmetric hydrogenation of various<br />

enamides. The influences of different reaction parameters<br />

are presented. For the first time high enantioselectivity (up<br />

to 95% ee) for such reactions is obtained in the presence of<br />

Eur. J. Org. Chem. 2006, 2912–2917 © 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.eurjoc.org 2915


FULL PAPER<br />

monodentate phosphanes. Best results were observed with<br />

aryl-substituted phosphepine ligands for reduction of N-(1phenylvinyl)acetamide,<br />

while alkyl-substituted phosphepine<br />

ligands yielded only low selectivity.<br />

Experimental Section<br />

All manipulations were performed under argon using standard<br />

Schlenk techniques. Diethyl ether and toluene were distilled from<br />

sodium benzophenone ketyl under argon. Methanol was distilled<br />

from Mg under argon. Dichloromethane was distilled from CaH2 under argon. The ligands 5 were synthesized according to our previously<br />

published protocols. [13] [Rh(cod) 2]BF4 (purchased from<br />

Fluka) was used without further purification.<br />

General Procedure for the Synthesis of Substituted N-(1-Phenylvinyl)acetamides<br />

6a–6e: To a stirred solution of methylmagnesium bromide<br />

(17.0 mmol, 3.0 mol/L diethyl ether, 6.0 mL) in diethyl ether<br />

(50 mL) at 0 °C a solution of the corresponding benzonitrile<br />

(17.0 mmol) in diethyl ether (20 mL) was added dropwise during a<br />

period of 30 minutes. After complete addition the solution was refluxed<br />

for eight hours. Within a few hours a yellow precipitate was<br />

formed. After refluxing the reaction mixture was cooled to 0 °C,<br />

and a solution of acetic anhydride (17.0 mmol) in diethyl ether<br />

(20 mL) was added care<strong>full</strong>y over 30 minutes. The reaction mixture<br />

was refluxed for eight hours. To the resulting suspension methanol<br />

was added at room temperature whilst stirring until all precipitates<br />

were dissolved (approximately 50 mL). The homogeneous solution<br />

was mixed with water/ethyl acetate (1:1, 100 mL). After phase separation<br />

the aqueous layer was extracted three times with ethyl acetate<br />

(50 mL). The combined organic layers were dried with MgSO4. After removing of the solvents the semi crystalline crude oil was<br />

purified by column chromatography (n-hexane/ethyl acetate, 1:1).<br />

Removal of the solvent yielded the crystalline products. [yields: (6a)<br />

1.51 g (55 %), (6b) 1.66 g (51 %), (6c) 1.50 g (46 %), (6d) 1.52 g<br />

(50%), (6e) 1.91 g (49 %)]<br />

General Procedure for the Synthesis of the Cyclic N-Acyl Enamide<br />

6f: A stirred solution of the corresponding ketone (30.3 mmol), hydroxylamine–hydrochloride<br />

(73 mmol) and pyridine (62.2 mmol) in<br />

ethanol (40 mL) was heated to 85 °C for a period of 16 hours. The<br />

solvent was removed, and the residue was dissolved in ethyl acetate/<br />

water. The organic phase was washed two times with water (20 mL)<br />

and dried with MgSO4. After removal of the solvents the product<br />

was recrystallized from toluene. The corresponding ketoximine<br />

(18.6 mmol) was solved in toluene (30 mL) under argon. To the<br />

stirred solution acetic acid anhydride (55.9 mmol), acetic acid<br />

(55.9 mmol), and Fe powder (37.3 mmol, Aldrich 325 mesh) were<br />

added and then heated to 70 °C for 4 hours. The mixture was filtered<br />

through a plug of celite after cooling to room temperature.<br />

Dichloromethane was added to the filtrate, followed by washing<br />

with 2.0 m NaOH (2 ×25 mL) at 0 °C. The separated organic phase<br />

was concentrated to half volume. The crystalline product was obtained<br />

after 12 hours at 0 °C. The crystals were filtered and dried<br />

in vacuo. [overall yield: 2.05 g (59 %)]<br />

General Procedure for the Catalytic Hydrogenation of Enamides: A<br />

solution of enamide (0.24 mmol) and 1.0 mL solvent was transferred<br />

via syringe into the secured autoclave. Because of the poor<br />

solubility in toluene at room temperature the solution was heated<br />

to 50 °C before it was transferred. The catalyst was generated in<br />

situ by mixing [Rh(cod) 2]BF4 (0.0024 mmol) and the corresponding<br />

4,5-dihydro-3H-dinaphthophosphepine ligands (0.005 mmol) in<br />

1.0 mL solvent for a period of 10 min, and afterwards, it was trans-<br />

2916<br />

S. Enthaler, B. Hagemann, K. Junge, G. Erre, M. Beller<br />

ferred via syringe into the autoclave. Then, the autoclave was<br />

charged with hydrogen and the mixture was stirred at the required<br />

temperature. After the predetermined time the hydrogen was released<br />

and the reaction mixture passed through a short plug of<br />

silica gel. The enantioselectivity and conversion were measured by<br />

GC or HPLC without further modifications.<br />

Acknowledgments<br />

We thank M. Heyken, S. Buchholz and Dr. C. Fischer (all Leibniz-<br />

Institut für Katalyse e. V. an der <strong>Universität</strong> <strong>Rostock</strong>) for excellent<br />

technical assistance and Prof. Dr. S. Gladiali for general discussions.<br />

Generous financial support from the state of Mecklenburg-Western<br />

Pomerania and the BMBF is grate<strong>full</strong>y acknowledged.<br />

[1] a) Y.-G. Zhou, P.-Y. Yang, X. W. Han, J. Org. Chem. 2005, 70,<br />

1679–1683; b) P. Harrison, G. Meek, Tetrahedron Lett. 2004,<br />

45, 9277–9280; c) S. A. Lawrence, Amines, Cambridge University<br />

Press, Cambridge, 2004; d) F. Fache, E. Schulz, M. L. Tommasino,<br />

M. Lemaire, Chem. Rev. 2000, 100, 2159–2231; e) H.-<br />

P. Jacqueline, Chiral Auxiliaries and Ligands in Asymmetric<br />

Synthesis, John Wiley & Sons, New York, 1995; f)M.Nógradi,<br />

Stereoselective Synthesis, Wiley-VCH, Weinheim, 2 nd ed., 1995;<br />

g) D. M. Tschaen, L. Abramson, D. Cai, R. Desmond, U.-H.<br />

Dolling, L. Frey, S. Karady, Y.-J. Shi, T. R. Verhoeven, J. Org.<br />

Chem. 1995, 60, 4324–4330; h) P. A. Chaloner, M. A. Esteruelas,<br />

F. Joó, L.A. Oro, Homogeneous Hydrogenation,<br />

Kluwer, Dordrecht, 1994; i) M. Kitamura, Y. Hsiao, M. Ohta,<br />

M. Tsukamoto, T. Ohta, H. Takaya, R. Noyori, J. Org. Chem.<br />

1994, 59, 297–310; j) H. P. Buser, B. Pugin, F. Spindler, M.<br />

Sutter, Tetrahedron 1991, 47, 5709–5716; k) R. Noyori, M.<br />

Ohta, Y. Hsiao, M. Kitamura, T. Ohta, H. Takaya, J. Am.<br />

Chem. Soc. 1986, 108, 7117–7119.<br />

[2] a) R. Noyori, Asymmetric Catalysis in Organic Synthesis, Wiley,<br />

New York, 1994; b) M. Beller, C. Bolm (Eds.), Transition Metals<br />

for Organic Synthesis, Wiley-VCH, Weinheim, 2 nd ed., 2004;<br />

c) E. N. Jacobsen, A. Pfaltz, H. Yamamoto (Eds.), Comprehensive<br />

Asymmetric Catalysis, Springer, Berlin, 1999.<br />

[3] H.-U. Blaser, B. Pugin, F. Spindler, J. Mol. Catal. A 2005, 231,<br />

1–20.<br />

[4] a) H. B. Kagan, N. Langlois, T. P. Dang, J. Organomet. Chem.<br />

1975, 90, 353–365; b) D. Sinou, H. B. Kagan, J. Organomet.<br />

Chem. 1976, 114, 325–337.<br />

[5] a) M. J. Burk, J. E. Feaster, W. A. Nugent, R. L. Harlow, J.<br />

Am. Chem. Soc. 1993, 115, 10125–12138; b) A. Miyashita, A.<br />

Yasuda, H. Takaya, K. Toriumi, T. Ito, T. Souchi, R. Noyori,<br />

J. Am. Chem. Soc. 1980, 102, 7932–7934; c) W. Li, X. Zhang,<br />

J. Org. Chem. 2000, 65, 5871–5874; d) W. Hu, M. Yan, C.-P.<br />

Lau, S. M. Yang, A. S. C. Chan, Tetrahedron Lett. 1999, 40,<br />

973–976; e) Z. Zhang, G. Zhu, Q. Jiang, D. Xiao, X. Zhang,<br />

J. Org. Chem. 1999, 64, 1774–1775; f) P. Dupau, P. Le Gendre,<br />

C. Bruneau, P. H. Dixneuf, Synlett 1999, 11, 1832–1834; g) G.<br />

Zhu, X. Zhang, J. Org. Chem. 1998, 63, 9590–9593; h) T. Morimoto,<br />

M. Chiba, K. Achiwa, Chem. Pharm. Bull. 1992, 40,<br />

2894–2896; i) M. J. Burk, Y. M. Wang, J. R. Lee, J. Am. Chem.<br />

Soc. 1996, 118, 5142–5143; j) F.-Y. Zhang, C.-C. Pai, A. S. C.<br />

Chan, J. Am. Chem. Soc. 1998, 120, 5808–5809; k) M. Hayashi,<br />

Y. Hashimoto, H. Takezaki, Y. Watanabe, K. Saigo, Tetrahedron:<br />

Asymmetry 1998, 9, 1863–1866.<br />

[6] F. Lagasse, H. B. Kagan, Chem. Pharm. Bull. 2000, 48, 315–<br />

324.<br />

[7] I. V. Komarov, A. Börner, Angew. Chem. 2001, 113, 1237–1240.<br />

[8] a) L. A. Arnold, R. Imobos, A. Manoli, A. H. M. de Vries, R.<br />

Naasz, B. Feringa, Tetrahedron 2000, 56, 2865–2878; b) M.<br />

van den Berg, A. J. Minnaard, E. P. Schudde, J. van Esch,<br />

A. H. M. de Vries, J. G. de Vries, B. L. Feringa, J. Am. Chem.<br />

Soc. 2000, 122, 11539–11540; c) A. J. Minnaard, M.<br />

www.eurjoc.org © 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Eur. J. Org. Chem. 2006, 2912–2917


Enantioselective Hydrogenation of Enamides in the Presence of Phosphanes FULL PAPER<br />

van den Berg, E. P. Schudde, J. van Esch, A. H. M. de Vries,<br />

J. G. de Vries, B. Feringa, Chim. Oggi 2001, 19, 12–13; d) D.<br />

Peña, A. J. Minnaard, J. G. de Vries, B. L. Feringa, J. Am.<br />

Chem. Soc. 2002, 124, 14552–14553; e) M. van den Berg, R. M.<br />

Haak, A. J. Minnaard, A. H. M. de Vries, J. G. de Vries, B. L.<br />

Feringa, Adv. Synth. Catal. 2002, 344, 1003–1007; f) M.<br />

van den Berg, A. J. Minnaard, R. M. Haak, M. Leeman, E. P.<br />

Schudde, A. Meetsma, B. L. Feringa, A. H. M. de Vries,<br />

C. E. P. Maljaars, C. E. Willans, D. Hyett, J. A. F. Boogers,<br />

H. J. W. Henderickx, J. G. de Vries, Adv. Synth. Catal. 2003,<br />

345, 308–323; g) D. Peña, A. J. Minnaard, A. H. M. de Vries,<br />

J. G. de Vries, B. L. Feringa, Org. Lett. 2003, 5, 475–478; h)<br />

X. Jiang, M. van den Berg, A. J. Minnaard, B. Feringa, J. G.<br />

de Vries, Tetrahedron: Asymmetry 2004, 15, 2223–2229; i) R.<br />

Eelkema, R. A. van Delden, B. L. Feringa, Angew. Chem. Int.<br />

Ed. 2004, 43, 5013–5016; j) A. Duursma, J.-G. Boiteau, L. Lefort,<br />

J. A. F. Boogers, A. H. M. de Vries, J. G. de Vries, A. J.<br />

Minnaard, B. L. Feringa, J. Org. Chem. 2004, 69, 8045–8052;<br />

k) H. Bernsmann, M. van den Berg, R. Hoen, A. J. Minnaard,<br />

G. Mehler, M. T. Reetz, J. G. de Vries, B. L. Feringa, J. Org.<br />

Chem. 2005, 70, 943–951.<br />

[9] a) M. T. Reetz, T. Sell, Tetrahedron Lett. 2000, 41, 6333–6336;<br />

b) M. T. Reetz, G. Mehler, Angew. Chem. Int. Ed. 2000, 39,<br />

3889–3891; c) M. T. Reetz, G. Mehler, G. Meiswinkel, T. Sell,<br />

Tetrahedron Lett. 2002, 43, 7941–7943; d) M. T. Reetz, G.<br />

Mehler, G. Meiswinkel, T. Sell, Angew. Chem. Int. Ed. 2003,<br />

42, 790–793; e) M. T. Reetz, G. Mehler, Tetrahedron Lett. 2003,<br />

44, 4593–4596; f) M. T. Reetz, G. Mehler, A. Meiswinkel, Tetrahedron:<br />

Asymmetry 2004, 15, 2165–2167; g) M. T. Reetz,<br />

X. G. Li, Tetrahedron 2004, 60, 9709–9714; h) M. T. Reetz, J.-<br />

A. Ma, R. Goddard, Angew. Chem. Int. Ed. 2005, 44, 412–414.<br />

[10] a) C. Claver, E. Fernandez, A. Gillon, K. Heslop, D. J. Hyett,<br />

A. Martorell, A. G. Orpen, P. G. Pringle, Chem. Commun.<br />

2000, 961–962; b) A. Martorell, R. Naasz, B. L. Feringa, P. G.<br />

Pringle, Tetrahedron: Asymmetry 2001, 12, 2497–2499.<br />

[11] A.-G. Hu, Y. Fu, J.-H. Xie, H. Zhou, L.-X. Wang, Q.-L. Zhou,<br />

Angew. Chem. 2002, 114, 2454–2456.<br />

[12] a) S. Gladiali, A. Dore, D. Fabbri, O. De Lucchi, M. Manassero,<br />

Tetrahedron: Asymmetry 1994, 5, 511–514; b) L. Charruault,<br />

V. Michelet, R. Taras, S. Gladiali, J.-P. Genet, Chem.<br />

Commun. 2004, 850–851; c) G. Zanoni, S. Gladiali, A. Marchetti,<br />

P. Piccinini, I. Tredici, G. Vidari, Angew. Chem. Int. Ed.<br />

2004, 43, 846–849.<br />

[13] a) Y. Chi, X. Zhang, Tetrahedron Lett. 2002, 43, 4849–4852; b)<br />

W. Tang, W. Wang, Y. Chi, X. Zhang, Angew. Chem. Int. Ed.<br />

2003, 42, 3506–3509; c) Y. Chi, Y.-G. Zhou, X. Zhang, J. Org.<br />

Chem. 2003, 68, 4120–4122; d) D. Xiao, X. Zhang, Angew.<br />

Chem. Int. Ed. 2001, 40, 3425–3428; e) D. Xiao, Z. Zhang, X.<br />

Zhang, Org. Lett. 1999, 1, 1679–1681.<br />

[14] a) K. Junge, G. Oehme, A. Monsees, T. Riermeier, U. Dingerdissen,<br />

M. Beller, Tetrahedron Lett. 2002, 43, 4977–4980; b) K.<br />

Junge, G. Oehme, A. Monsees, T. Riermeier, U. Dingerdissen,<br />

M. Beller, J. Organomet. Chem. 2003, 675, 91–96; c) K. Junge,<br />

B. Hagemann, S. Enthaler, A. Spannenberg, M. Michalik, G.<br />

Oehme, A. Monsees, T. Riermeier, M. Beller, Tetrahedron:<br />

Asymmetry 2004, 15, 2621–2631; d) K. Junge, B. Hagemann,<br />

S. Enthaler, G. Oehme, M. Michalik, A. Monsees, T. Riermeier,<br />

U. Dingerdissen, M. Beller, Angew. Chem. 2004, 116, 5176–<br />

5179; Angew. Chem. Int. Ed. 2004, 43, 5066–5069.<br />

[15] a) G. Zanoni, S. Gladiali, A. Marchetti, P. Piccinini, I. Tredici,<br />

G. Vidari, Angew. Chem. 2004, 116, 864–867; b) R. P. Wurz,<br />

G. C. Fu, J. Am. Chem. Soc. 2005, 127, 12234–12235; c) L.<br />

Charruault, V. Michelet, R. Tarras, S. Gladiali, J.-P. Genêt,<br />

Chem. Commun. 2004, 850–851; d) E. Vedejs, O. Daugulis, S. T.<br />

Diver, J. Org. Chem. 1996, 61, 430–431.<br />

[16] M. T. Reetz, G. Mehler, A. Meiswinkel, Tetrahedron: Asymmetry<br />

2004, 15, 2165–2167.<br />

[17] M. J. Burk, G. Casy, N. B. Johnson, J. Org. Chem. 1998, 63,<br />

6084–6085.<br />

[18] Institute of Automation (IAT), Richard Wagner Str. 31/ H. 8,<br />

18119 <strong>Rostock</strong>-Warnemünde, Germany.<br />

[19] X. Jia, X. Li, L. Xu, Q. Shi, X. Yao, A. S. C. Chan, J. Org.<br />

Chem. 2003, 68, 4539–4541.<br />

[20] MonoPhos is commercially available by Strem.<br />

[21] a) M. T. Reetz, X. Li, Angew. Chem. Int. Ed. 2005, 44, 2959–<br />

2962; b) M. T. Reetz, X. Li, Angew. Chem. Int. Ed. 2005, 44,<br />

2962–2964.<br />

[22] a) A. Duursma, R. Hoen, J. Schuppan, R. Hulst, A. J. Minnaard,<br />

B. L. Feringa, Org. Lett. 2003, 5, 3111–3113; b) A. Duursma,<br />

D. Peña, A. J. Minnaard, B. L. Feringa, Tetrahedron:<br />

Asymmetry 2005, 16, 1901–1904; c) D. Peña, A. J. Minnaard,<br />

J. A. F. Boogers, A. H. M. de Vries, B. L. Feringa, Org. Biomol.<br />

Chem. 2003, 1, 1087–1089.<br />

[23] a) C. Monti, C. Gennari, U. Piarulli, J. G. de Vries, A. H. M.<br />

de Vries, L. Lefort, Chem. Eur. J. 2005, 11, 6701–6717; b) K.<br />

Ding, H. Du, Y. Yuan, J. Long, Chem. Eur. J. 2004, 10, 2872–<br />

2884.<br />

Received: January 13, 2006<br />

Published Online: April 26, 2006<br />

Eur. J. Org. Chem. 2006, 2912–2917 © 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.eurjoc.org 2917


5. Application of monodentate phosphine ligands in asymmetric hydrogenation 76<br />

5.2. Development of Practical Rhodium Phosphine Catalysts for the<br />

Hydrogenation of β–Dehydroamino Acid Derivates<br />

Stephan Enthaler, Giulia Erre, Kathrin Junge, Jens Holz, Armin Börner, Elisabetta Alberico,<br />

Ilenia Nieddu, Serafino Gladiali*, and Matthias Beller*, Org. Process Res. Dev. 2007, 11,<br />

568–577.<br />

Contributions<br />

Dr. Jens Holz carried out preliminary experiments with BINEPINE–ligands which are not<br />

stated in the article. Investigations with different metal precursors and defined complexes for<br />

instance catalysis with [Rh(6a)2(nbd)]SO3CF3 and [Rh(nbd)2]BF4 were realized by the group<br />

of Prof. Gladiali. Furthermore, ligands (R)–6a and 6e were synthesized by Dr. B. Hagemann.


Organic Process Research & Development 2007, 11, 568−577<br />

Development of Practical Rhodium Phosphine Catalysts for the Hydrogenation<br />

of �-Dehydroamino Acid Derivatives<br />

Stephan Enthaler, † Giulia Erre, † Kathrin Junge, † Jens Holz, † Armin Börner, †,‡ Elisabetta Alberico, § Ilenia Nieddu, |<br />

Serafino Gladiali,* ,| and Matthias Beller* ,†<br />

Leibniz Institut für Katalyse an der UniVersität <strong>Rostock</strong> e.V., Albert-Einstein Strasse 29a, 18059 <strong>Rostock</strong>, Germany,<br />

Institut für Chemie der UniVersität <strong>Rostock</strong>, Albert-Einstein Strasse 3a, 18059 <strong>Rostock</strong>, Germany, Istituto di Chimica<br />

Biomolecolare - CNR, traV. La Crucca 3, 07040 Sassari, Italy, and Dipartimento di Chimica, UniVersità di Sassari,<br />

Via Vienna 2, 07100 Sassari, Italy<br />

Abstract:<br />

The rhodium-catalyzed asymmetric hydrogenation of various<br />

�-dehydroamino acid derivatives to give optically active �-amino<br />

acids has been examined. Chiral monodentate 4,5-dihydro-3Hdinaphthophosphepines,<br />

which are easily tuned and accessible<br />

in a multi-10-g scale, have been used as ligands. The enantioselectivity<br />

is largely dependent on the nature of the substituent<br />

at the phosphorous atom and on the structure of the substrate.<br />

Applying optimized conditions up to 94% ee was achieved.<br />

Introduction<br />

The discovery of new effective drugs is an important<br />

challenge for industrial and academic research. During the<br />

last decades an intense area of research in medicinal<br />

chemistry was the development of peptide-based therapeutics,<br />

mainly constructed by R-amino acids. More recently, significant<br />

attention was also directed to non-natural �-amino<br />

acids, which are interesting building blocks for the synthesis<br />

of biologically active compounds such as �-lactam antibiotics,<br />

the taxol derivatives, and �-peptides (Scheme 1). 1 In<br />

view of the growing demand of chiral �-amino acids, an<br />

increasing number of synthetic methods have been established<br />

for their preparation such as homologation of R-amino<br />

acids, conjugate addition of amines to carbonyl compounds,<br />

Mannich reaction, hydrogenation etc. 1c,2 Within these methodologies,<br />

asymmetric catalytic hydrogenation constitutes the<br />

most attractive and versatile technology as to industrial<br />

applications due to the remarkable improvements achieved<br />

in the past few years in the asymmetric hydrogenation of<br />

�-dehydroamino acid derivatives. Good-to-excellent enan-<br />

* To whom correspondence should be addressed. E-mail: matthias.beller@<br />

catalysis.de. Fax: 49 381 12815000. Telephone: 49 381 1281113.<br />

† Leibniz Institut für Katalyse e.V an der <strong>Universität</strong> <strong>Rostock</strong>.<br />

‡ Institut für Chemie der <strong>Universität</strong> <strong>Rostock</strong>.<br />

§ Istituto di Chimica Biomolecolare - CNR.<br />

| Dipartimento di Chimica, Università di Sassari.<br />

(1) (a) Guénard, D.; Guritte-Voegelein, R.; Potier, P. Acc. Chem. Res. 1993,<br />

26, 160-167. (b) Cole, D. C. Tetrahedron 1994, 50, 9517-9582. (c) Juaristi,<br />

E.; Soloshonok, V. EnantioselectiVe Synthesis of �-amino acids, 2nd ed.;<br />

Wiley-Interscience: New Jersey, 2005. (d) Lui, M.; Sibi, M. P. Tetrahedron<br />

2002, 58, 7991-8035. (e) Drexler, H.-J.; You, J.; Zhang, S.; Fischer, C.;<br />

Baumann, W.; Spannenberg, A.; Heller, D. Org. Process Res. DeV. 2003,<br />

7, 355-361. (f) Ma, J. Angew. Chem. 2003, 115, 4426-4435. (g) Cheng,<br />

R. P.; Gellman, S. H.; DeGrado, W. F. Chem. ReV. 2001, 101, 3219-3232.<br />

(h) Kubota, D.; Ishikawa, M.; Yamamoto, M.; Murakami, S.; Hachisu, M.;<br />

Katano, K.; Ajito, K., Bioorg. Med. Chem. 2006, 7, 2089-2108.<br />

(2) (a) Gellman, S. H., Acc. Chem. Res. 1998, 31, 173-180. (b) Seebach, D.;<br />

Matthews, J. L. Chem. Commun. 1997, 2015-2022.<br />

tioselectivities have been obtained in this reaction by<br />

applying rhodium- or ruthenium-catalysts with chiral diphosphines<br />

such as MeDuPhos, catASium M, BINAP, BINAPO,<br />

BICP, TunaPhos, FerroTane, JosiPhos, DIOP, BPPM, P-Phos<br />

and others. 3<br />

A current trend in asymmetric catalysis is to switch from<br />

chiral bidentate to chiral monodentate phosphines because<br />

the latter are more easily accessible and tuneable than the<br />

bidentate counterparts. 4,5 Seminal contributions in this field<br />

have come from Feringa and de Vries et al. 6 (phosphoramidites),<br />

Reetz et al. 7 (phosphites), and Pringle and Claver<br />

et al. 8 (phosphonites). Furthermore, excellent enantioselectivities<br />

were reported by Zhou and co-workers applying<br />

monodentate spiro phosphoramidites (SIPHOS). 9,10<br />

Following the first report by Gladiali 11 and parallel to the<br />

work of Zhang, 12 we have focused in recent years on different<br />

monodentate phosphines based on a 4,5-dihydro-3H-dinaphtho[2,1-c;1′,2′-e]phosphepine<br />

framework (5 and 6a-j)<br />

(3) (a) Vineyard, B. D.; Knowles, W. S.; Sabacky, M. J.; Bachman, G. L.;<br />

Weinkauff, D. J. J. Am. Chem. Soc. 1977, 99, 5947-5952. (b) Achiwa, K.;<br />

Soga, T. Tetrahedron Lett. 1978, 13, 1119-1120. (c) Lubell, W. D.;<br />

Kitamura, M.; Noyori, R. Tetrahedron: Asymmetry 1991, 2, 543-554. (d)<br />

Zhu, G.; Chen, Z.; Zhang, X. J. Org. Chem. 1999, 64, 6907-6910. (e)<br />

Heller, D.; Holz, J.; Drexler, H.-J.; Lang, J.; Drauz, K.; Krimmer, H.-P.;<br />

Börner, A. J. Org. Chem. 2001, 66, 6816-6817. (f) Yasutake, M.; Gridnev,<br />

I. D.; Higashi, N.; Imamoto, T. Org. Lett. 2001, 3, 1701-1704. (g) Heller,<br />

D.; Drexler, H.-J.; You, J.; Baumann, W.; Drauz, K.; Krimmer, H.-P.;<br />

Börner, A. Chem. Eur. J. 2002, 8, 5196-5203. (h) Zhou, Y.-G.; Tang, W.;<br />

Wang, W.-B.; Li, W.; Zhang, X. J. Am. Chem. Soc. 2002, 124, 4952-<br />

4953. (i) Yang, W.; Wu, S.; Zhang, X. J. Am. Chem. Soc. 2003, 125, 9570-<br />

9571. (j) Komarov, I. V.; Monsees, A.; Spannenberg, A.; Baumann, W.;<br />

Schmidt, U.; Fischer, C.; Börner, A. Eur. J. Org. Chem. 2003, 138-150.<br />

(k) You, J.; Drexler, H.-J.; Zhang, S.; Fischer, C.; Heller, D. Angew. Chem.<br />

2003, 115, 942-945. (l) Holz, J.; Monsees, A.; Jiao, H.; You, J.; Komarov,<br />

I. V.; Fischer, C.; Drauz, K.; Börner, A. J. Org. Chem. 2003, 68, 1701-<br />

1707. (m) Wu, J.; Chen, X.; Guo, R.; Yeung, C.-h.; Chan, A. S. C. J. Org.<br />

Chem. 2003, 68, 2490-2493. (n) Henschke, J. P.; Burk, M. J.; Malan, C.<br />

G.; Herzberg, D.; Peterson, J. A.; Wildsmith, A. J.; Cobley, C. J.; Casy, G.<br />

AdV. Synth. Catal. 2003, 345, 300-307. (o) Robinson, A. J.; Lim, C. Y.;<br />

He, L.; Ma, P.; Li, H.-Y. J. Org. Chem. 2001, 66, 4141-4147. (p) Robinson,<br />

A. J.; Stanislawski, P.; Mulholland, D.; He, L.; Li, H.-Y. J. Org. Chem.<br />

2001, 66, 4148-4152. (q) Wu, H.-P.; Hoge, G. Org. Lett. 2004, 6, 3645-<br />

3647. (r) Hsiao, Y.; Rivera, N. R.; Rosner, T.; Krska, S. W.; Njolito, E.;<br />

Wang, F.; Sun, Y.; Armstrong, J. D., III; Grabowski, E. J. J.; Tillyer, R.<br />

D.; Spindler, F.; Malan, C. J. Am. Chem. Soc. 2004, 126, 9918-9919. (s)<br />

Dai. Q.; Yang, W.; Zhang, X. Org. Lett. 2005, 7, 5343-5345. (t) Hansen,<br />

K. B.; Rosner, T.; Kubryk, M.; Dormer, P. G.; Armstrong, J. D., III. Org.<br />

Lett. 2005, 7, 4935-4938. (u) Drexler, H.-J.; Baumann, W.; Schmidt, T.;<br />

Zhang, S.; Sun, A.; Spannenberg, A.; Fischer, C.; Buschmann, H.; Heller,<br />

D. Angew. Chem. 2005, 117, 1208-1212. (v) Holz, J.; Zayas, O.; Jiao, H.;<br />

Baumann, W.; Spannenberg, A.; Monsees, A.; Riermeier, T. H.; Almena,<br />

J.; Kadyrov, R.; Börner, A. Chem. Eur. J. 2006, 12, 5001-5013.<br />

(4) Lagasse, F.; Kagan, H. B. Chem. Pharm. Bull. 2000, 48, 315-324.<br />

(5) Komarov, I. V.; Börner, A. Angew. Chem. 2001, 113, 1237-1240.<br />

568 • Vol. 11, No. 3, 2007 / Organic Process Research & Development 10.1021/op0602270 CCC: $37.00 © 2007 American Chemical Society<br />

Published on Web 12/30/2006


Scheme 1. Selection of biologically active compounds containing �-amino acid units<br />

(Scheme 3). 13 The first catalytic application of such a ligand<br />

(Ph-BINEPINE; 6a) in the Rh-catalyzed asymmetric hydroformylation<br />

of styrene was reported by one of us (S.G.) in<br />

the early 1990s. 11a In the following years, the <strong>Rostock</strong> group<br />

has expanded the structural diversity of the BINEPINE ligand<br />

family 6 into a library of ligands which has been screened<br />

with remarkable success (ee up to 95%) in the asymmetric<br />

(6) (a) Arnold, L. A.; Imobos, R.; Manoli, A.; de Vries, A. H. M.; Naasz, R.;<br />

Feringa, B. Tetrahedron 2000, 56, 2865-2878. (b) van den Berg, M.;<br />

Minnaard, A. J.; Schudde, E. P.; van Esch, J.; de Vries, A. H. M.; de Vries,<br />

J. G.; Feringa, B. L. J. Am. Chem. Soc. 2000, 122, 11539-11540. (c)<br />

Minnaard, A. J.; van den Berg, M.; Schudde, E. P.; van Esch, J.; de Vries,<br />

A. H. M.; de Vries, J. G.; Feringa, B. Chim. Oggi 2001, 19, 12-13. (d)<br />

Peña, D.; Minnaard, A. J.; de Vries, J. G.; Feringa, B. L. J. Am. Chem.<br />

Soc. 2002, 124, 14552-14553. (e) van den Berg, M.; Haak, R. M.;<br />

Minnaard, A. J.; de Vries, A. H. M.; de Vries, J. G.; Feringa, B. L.; AdV.<br />

Synth. Catal. 2002, 344, 1003-1007. (f) van den Berg, M.; Minnaard, A.<br />

J.; Haak, R. M.; Leeman, M.; Schudde, E. P.; Meetsma, A.; Feringa, B. L.;<br />

de Vries, A. H. M.; Maljaars, C. E. P.; Willans, C. E.; Hyett, D.; Boogers,<br />

J. A. F.; Henderickx, H. J. W.; de Vries, J. G. AdV. Synth. Catal. 2003,<br />

345, 308-323. (g) Peña, D.; Minnaard, A. J.; de Vries, A. H. M.; de Vries,<br />

J. G.; Feringa, B. L.; Org. Lett. 2003, 5, 475-478. (h) Jiang, X.; van den<br />

Berg, M.; Minnaard, A. J.; Feringa, B.; de Vries, J. G. Tetrahedron:<br />

Asymmetry 2004, 15, 2223-2229. (i) Eelkema, R.; van Delden, R. A.;<br />

Feringa, B. L. Angew. Chem., Int. Ed. 2004, 43, 5013-5016. (j) Duursma,<br />

A.; Boiteau, J.-G.; Lefort, L.; Boogers, J. A. F.; de Vries, A. H. M.; de<br />

Vries, J. G.; Minnaard, A. J.; Feringa, B. L. J. Org. Chem. 2004, 69, 8045-<br />

8052. (k) Lefort, L.; Boogers, J. A. F.; de Vries, A. H. M.; de Vries, J. G.<br />

Org. Lett. 2004, 6, 1733-1735. (l) Bernsmann, H.; van den Berg, M.; Hoen,<br />

R.; Minnaard, A. J.; Mehler, G.; Reetz, M. T.; de Vries, J. G.; Feringa, B.<br />

L. J. Org. Chem. 2005, 70, 943-951.<br />

(7) (a) Reetz, M. T.; Sell, T. Tetrahedron Lett. 2000, 41, 6333-6336. (b) Reetz,<br />

M. T.; Mehler, G. Angew. Chem., Int. Ed. 2000, 39, 3889-3891. (c) Reetz,<br />

M. T.; Mehler, G.; Meiswinkel, G.; Sell, T. Tetrahedron Lett. 2002, 43,<br />

7941-7493. (d) Reetz, M. T.; Mehler, G.; Meiswinkel, G.; Sell, T. Angew.<br />

Chem., Int. Ed. 2003, 42, 790-793. (e) Reetz, M. T.; Mehler, G. Tetrahedron<br />

Lett. 2003, 44, 4593-4596. (f) Reetz, M. T.; Mehler, G.; Meiswinkel, A.<br />

Tetrahedron: Asymmetry 2004, 15, 2165-2167. (g) Reetz, M. T.; Li, X.<br />

G. Tetrahedron 2004, 60, 9709-9714. (h) Reetz, M. T.; Ma, J.-A.; Goddard,<br />

R. Angew. Chem., Int. Ed. 2005, 44, 412-414. (i) Reetz, M. T.; Fu, Y.;<br />

Meiswinkel, A. Angew. Chem. 2006, 118, 1440-1443.<br />

(8) (a) Claver, C.; Fernandez, E.; Gillon, A.; Heslop, K.; Hyett, D. J.; Martorell,<br />

A.; Orpen, A. G.; Pringle, P. G. Chem. Commun. 2000, 961-962. (b)<br />

Martorell, A.; Naasz, R.; Feringa, B. L.; Pringle, P. G. Tetrahedron:<br />

Asymmetry 2001, 12, 2497-2499.<br />

(9) (a) Hu, A.-G.; Fu, Y.; Xie, J.-H.; Zhou, H.; Wang, L.-X.; Zhou, Q.-L. Angew.<br />

Chem. 2002, 114, 2454-2456. (b) Fu, Y.; Guo, X.-X.; Zhu, S.-F.; Hu, A.-<br />

G.; Xie, J.-H.; Zhou, Q.-L. J. Org. Chem. 2004, 69, 4648-4655. (c) Fu,<br />

Y.; Hou, G.-H.; Xie, J.-H.; Xing, L.; Wang, L.-X.; Zhou, Q.-L. J. Org.<br />

Chem. 2004, 69, 8157-8160.<br />

(10) For other monodentate system see also: (a) Jerphagnon, T.; Renaud, L.-L.;<br />

Demonchauc, P.; Ferreira, A.; Bruneau, C. AdV. Synth. Catal. 2004, 346,<br />

33-36. (b) Bilenko, V.; Spannenberg, A.; Baumann, W.; Komarov, I.;<br />

Börner, A. Tetrahedron: Asymmetry 2006, 17, 2082-2087.<br />

hydrogenation with Rh- and Ru-catalysts of R-amino acid<br />

precursors, dimethyl itaconate, enamides and �-ketoesters. 13<br />

In the meantime, the utility of Ph-BINEPINE 6a has been<br />

demonstrated in a range of asymmetric reactions catalyzed<br />

by different transition metals such as the Pd-catalyzed<br />

umpoled-allylation of aldehydes and the Pt-catalyzed alkoxycyclization<br />

of 1,5-enynes. The ligand itself without any metal<br />

is an efficient chiral catalyst for the enantioselective acylation<br />

of diols, and for [3 + 2] cycloadditions and [4 + 2]<br />

annulations (Scheme 2). 11d,e,14 Quite recently excellent enantioselectivities<br />

have been scored in the Rh-catalyzed transfer<br />

hydrogenation of R-amino acid precursors and itaconic acid<br />

derivatives using formic acid as H-donor. Under these<br />

conditions �-dehydroamino acids derivatives have also been<br />

success<strong>full</strong>y hydrogenated, albeit in modest stereoselectivity. 11f<br />

Pursuing our ongoing research in hydrogenation chemistry,<br />

we report herein on the Rh-catalyzed asymmetric<br />

hydrogenation of �-dehydroamino acid derivatives. A multi-<br />

10-g scale synthesis of binaphthophosphepines is described,<br />

which allowed these ligands to be commercialized last year. 15<br />

(11) (a) Gladiali, S.; Dore, A.; Fabbri, D.; De Lucchi, O.; Manassero, M.<br />

Tetrahedron: Asymmmetry 1994, 5, 511-514. (b) Gladiali, S.; Dore, A.;<br />

Fabbri, D.; De Lucchi, O.; Valle, G. J. Org. Chem. 1994, 59, 6363-6371.<br />

(c) Gladiali, S.; Frabbi, D. Chem. Ber./ Rec. TraV. Chim. Pays Bas. 1997,<br />

130, 543-554. (d) Charruault, L.; Michelet, V.; Taras, R.; Gladiali, S.;<br />

Genêt, J.-P. Chem. Commun. 2004, 850-851. (e) Zanoni, G.; Gladiali, S.;<br />

Marchetti, A.; Piccinini, P.; Tredici, I.; Vidari, G. Angew. Chem., Int. Ed.<br />

2004, 43, 846-849. (f) Alberico, E.; Nieddu, I.; Taras, R.; Gladiali, S. HelV.<br />

Chim. Acta 2006, 89, 1716-1729.<br />

(12) (a) Chi, Y.; Zhang, X. Tetrahedron Lett. 2002, 43, 4849-4852. (b) Tang,<br />

W.; Wang, W.; Chi, Y.; Zhang, X. Angew. Chem., Int. Ed. 2003, 42, 3506-<br />

3509. (c) Chi, Y.; Zhou, Y.-G.; Zhang, X. J. Org. Chem. 2003, 68, 4120-<br />

4122. (d) Xiao, D.; Zhang, X. Angew. Chem., Int. Ed. 2001, 40, 3425-<br />

3428. (e) Xiao, D.; Zhang, Z.; Zhang, X. Org. Lett. 1999, 1, 1679-1681.<br />

(13) (a) Junge, K.; Oehme, G.; Monsees, A.; Riermeier, T.; Dingerdissen, U.;<br />

Beller, M. Tetrahedron Lett. 2002, 43, 4977-4980. (b) Junge, K.; Oehme,<br />

G.; Monsees, A.; Riermeier, T.; Dingerdissen, U.; Beller, M. J. Organomet.<br />

Chem. 2003, 675, 91-96. (c) Junge, K.; Hagemann, B.; Enthaler, S.;<br />

Spannenberg, A.; Michalik, M.; Oehme, G.; Monsees, A.; Riermeier, T.;<br />

Beller, M. Tetrahedron: Asymmetry 2004, 15, 2621-2631. (d) Junge, K.;<br />

Hagemann, B.; Enthaler, S.; Oehme, G.; Michalik, M.; Monsees, A.;<br />

Riermeier, T.; Dingerdissen, U.; Beller, M. Angew. Chem. 2004, 116, 5176-<br />

5179; Angew. Chem., Int. Ed. 2004, 43, 5066-5069. (e) Hagemann, B.;<br />

Junge, K.; Enthaler, S.; Michalik, M.; Riermeier, T.; Monsees, A.; Beller,<br />

M. AdV. Synth. Catal. 2005, 347, 1978-1986. (f) Enthaler, S.; Hagemann,<br />

B.; Junge, K.; Erre, G.; Beller, M. Eur. J. Org. Chem. 2006, 2912-2917.<br />

(14) (a) Wurz, R. P.; Fu, G. C. J. Am. Chem. Soc. 2005, 127, 12234-12235. (b)<br />

Vedejs, E.; Daugulis, O.; Diver, S. T. J. Org. Chem. 1996, 61, 430-431.<br />

(c) Wilson, J. E.; Fu, G. C. Angew. Chem., Int. Ed. 2006, 45, 1426-1429.<br />

Vol. 11, No. 3, 2007 / Organic Process Research & Development • 569


Scheme 2. Applications of ligand class 6 in asymmetric synthesis<br />

Results and Discussion<br />

The synthesis of binaphthophosphepines starts with diesterification<br />

of enantiomerically pure 2,2′-binaphthol 16 (98%<br />

ee) with trifluoromethanesulfonic acid anhydride in the<br />

presence of pyridine (Scheme 3). The corresponding diester<br />

was obtained in 99% yield, and subsequent nickel-catalyzed<br />

Kumada coupling with methyl magnesium bromide led to<br />

2,2′-dimethylbinaphthyl in 95% yield. 13,17 Two different<br />

synthetic strategies were established to obtain 4,5-dihydro-<br />

3H-dinaphtho[2,1-c;1′,2′-e]phosphepine ligands 6. On the one<br />

hand, double metalation of 2,2′-dimethylbinaphthyl with<br />

n-butyl lithium in the presence of TMEDA (N,N,N′,N′tetramethylethylenediamine)<br />

followed by quenching with<br />

commercially available dichlorophosphines gives ligands 6a<br />

(P-phenyl) and 6i (P-tert-butyl) in 60-83% yield. Both<br />

ligands have been synthesized on >10-g scale. In the second<br />

procedure the dilithiated dimethyl binaphthalene is quenched<br />

with diethylaminodichlorophosphine, giving the phosphepine<br />

5 which, upon treatment with gaseous HCl, is converted into<br />

4-chloro-4,5-dihydro-3H-dinaphtho[2,1-c;1′,2′-e]phosphepine<br />

in 80% yield. This enantiomerically pure chlorophosphine<br />

is easily coupled with various Grignard or lithium<br />

(15) The ligands Ph-BINEPINE (6a) and t-Bu-BINEPINE (6i) are commercially<br />

available by Degussa DHC (Degussa Homogeneous Catalysis, Rodenbacher<br />

Chaussee 4, building 097, 63457 Haunau (Wolfgang), Germany, (www.creavis.com/site_dhc/de/default.cfm))<br />

with the product names catASium<br />

KPh (6a) and catASium KtB (6i).<br />

(16) Optically pure 2,2′-binaphthol is available on large scale from RCA (Reuter<br />

Chemische Apparatebau KG), Engesserstr. 4, 79108 Freiburg, Germany.<br />

(17) The synthesis of 2,2′-bistriflate-1,1′-binaphthyl (1.2 mol, 615 g, 92%) and<br />

2,2′-dimethyl-1,1′-binaphthyl (0.74 mol, 201 g, 96%) was carried out in<br />

large scale by the group of Zhang (See ref 11b).<br />

570 • Vol. 11, No. 3, 2007 / Organic Process Research & Development<br />

reagents to give a broad selection of ligand 6. The limited<br />

number of commercially available dichlorophosphines and<br />

the large diversity of Grignard compounds make the access<br />

through 4-chloro-4,5-dihydro-3H-dinaphtho[2,1-c;1′,2′-e]phosphepine<br />

the route of choice to a library of ligands 6.<br />

The syntheses of �-acetamido acrylates 7-11 and 14 were<br />

carried out according to literature protocols by reaction of<br />

the corresponding �-ketoester with NH4OAc followed by<br />

acylation of the �-amino acrylate intermediate. 3c,d E/Z-<br />

Isomers were separated by crystallization and column chromatography.<br />

Ethyl 3-acetamido-3-phenyl-2-propenoate (12)<br />

and methyl 3-benzamido butenoate (13) were synthesized<br />

by reaction of the corresponding �-amino acrylate with acyl<br />

chloride or benzoyl chloride. 3l<br />

To compare the behaviour of the Z- and E-isomers, initial<br />

catalytic runs were carried out separately on the (Z)- and<br />

(E)-methyl 3-acetamido butenoates (Z-7 and E-7, respectively)<br />

as substrates and 4-phenyl-4,5-dihydro-3H-dinaphtho-<br />

[2,1-c;1′,2′-e]phosphepine (6a) as our standard ligand. 13<br />

Typically, we used an in situ precatalytic mixture of 1<br />

mol % [Rh(cod)2]BF4 and 2.1 mol % of the corresponding<br />

ligand. All hydrogenation reactions were carried out in an<br />

8-fold parallel reactor array with a reactor volume of 3.0<br />

mL. 18<br />

We first focused our attention on the influence of the<br />

solvent and the hydrogen pressure. The first set of reactions<br />

was run at constant concentration in toluene, dichloromethane,<br />

methanol, ethanol, 19 and 2-propanol at three<br />

(18) Institute of Automation (IAT), Richard Wagner Str. 31/ H. 8, 18119 <strong>Rostock</strong>-<br />

Warnemünde, Germany.


Scheme 3. Synthetic approach to 4,5-dihydro-3H-dinaphtho[2,1-c;1′,2′-e]phosphepine ligands 6<br />

different pressures (2.5, 10.0, and 50.0 bar). Selected results<br />

are presented in Figure 1.<br />

In the case of hydrogenation of E-7 the best enantioselectivity<br />

was consistently achieved at the lowest pressure (2.5<br />

bar) and ranged between 42-79% ee, depending on the<br />

solvent. After 24 h the reaction was complete only in<br />

2-propanol, whereas lower conversions were observed in<br />

ethanol, methanol, and toluene. 20 No reaction at all took place<br />

in dichloromethane. The stereoselectivity decreased upon<br />

increasing the pressure of hydrogen to 10.0 bar or 50.0 bar.<br />

From these results 2-propanol and 2.5 bar were selected as<br />

solvent and pressure of choice, respectively, for the hydrogenation<br />

of the E-isomer (conversion: >99%; enantioselectivity:<br />

79%). The hydrogenation of Z-7 resulted in higher<br />

enantioselectivities in all the solvents (up to 92% ee).<br />

Notably, compared to E-7 the configuration of the prevailing<br />

enantiomer was always opposite. Furthermore, in ethanol and<br />

in methanol the enantioselectivity/pressure trend was reversed,<br />

the top value (92% ee) having been reached at the<br />

highest pressure (50.0 bar) compared to 86% ee at 2.5 bar.<br />

Full conversions were obtained in all the experiments, except<br />

in toluene (37%) and in dichloromethane (45%). These<br />

(19) The hydrogenation performed in ethanol and 2-propanol led exclusively to<br />

product 7. No transesterification was observed.<br />

(20) In the case of toluene we assume an inhibition of the catalyst by the solvent,<br />

due to the formation of catalytical inactive rhodium-arene complexes.<br />

Heller, D.; Drexler, H.-J.; Spannenberg, A.; Heller, B.; You, J.; Baumann,<br />

W. Angew. Chem., Int. Ed. 2002, 41, 777-780.<br />

results indicated ethanol and methanol as the best solvents<br />

and 50.0 bar as the pressure of choice for the hydrogenation<br />

of the Z-isomer (conversion: >99%; enantioselectivity:<br />

92%).<br />

No incorporation of deuterium in the product was noticed<br />

upon running the reaction in methanol-d4. This is in line with<br />

the absence in the reduction process of any interference of<br />

transfer hydrogenation as well as of protonolysis of the<br />

Rh-C bond of an alkyl rhodium intermediate. 21 Two facets<br />

deserving mention are (1) the Z-isomer gives higher ee’s than<br />

the E-isomer, whereas in most cases the opposite behaviour<br />

has been reported 3 and (2) the configuration switches<br />

depending on the different geometry of the double bond,<br />

which has been scarcely reported. 3c,22<br />

Next, we investigated the influence of temperature on<br />

enantioselectivity and conversion as shown in Figure 2. 23<br />

Applying our model ligand 6a we found a pronounced<br />

negative effect on the enantioselectivity at higher temperatures<br />

for both isomers. Interestingly, the loss of enantioselectivity<br />

for the hydrogenation of E-7 is higher than for Z-7<br />

(10 °C: E-7 79% ee (R) and Z-7 88% ee (S); 90 °C: E-7<br />

rac and Z-7 60% ee (S)).<br />

(21) Tsukamoto, M.; Yoshimura, M.; Tsuda, K.; Kitamura, M. Tetrahedron 2006,<br />

62, 5448-5453.<br />

(22) Hu, X.-P.; Zheng, Z. Org. Lett. 2005, 7, 419-422.<br />

(23) Jerphagnon, T.; Renaud, J.-L.; Demonchaux, P.; Ferreira, A.; Bruneau, C.<br />

Tetrahedron: Asymmetry 2003, 14, 1973-1977.<br />

Vol. 11, No. 3, 2007 / Organic Process Research & Development • 571


Figure 1. Solvent and pressure variation. Reactions were<br />

carried out at 10 °C for 24 h with 0.0024 mmol [Rh(cod)2]BF4,<br />

0.005 mmol ligand 6a and 0.24 mmol substrate in 2.0 mL of<br />

solvent. Conversions and ee’s were determined by GC (50 m<br />

Chiraldex �-PM, 130 °C, (S)-7a 15.1 min, (R)-7a 16.4 min). The<br />

absolute configuration was determined by comparison with<br />

reported data. 3l<br />

For estimating the feasible enantioselectivity at lower<br />

temperature we analysed the corresponding Eyring plot 24 for<br />

both isomers. As a consequence we considered 10 °C as<br />

temperature of choice, because of good enantioselectivity<br />

and also acceptable reaction times.<br />

The original protocol used for the synthesis of methyl<br />

3-acetamido butenoate (7) gives a mixture of E- and<br />

Z-isomers whose composition depends on the reaction<br />

conditions. 1f Although separation of the isomers by fractional<br />

(24) Buschmann, H.; Scharf, H.-D.; Hoffmann, N.; Esser, P. Angew. Chem. 1991,<br />

103, 480-518.<br />

572 • Vol. 11, No. 3, 2007 / Organic Process Research & Development<br />

Figure 2. Dependency of enantioselectivity versus temperature.<br />

Reactions were carried out at corresponding temperature for<br />

1-24 h with 0.0024 mmol [Rh(cod)2]BF4, 0.005 mmol ligand<br />

6a and 0.24 mmol substrate in 2.0 mL of 2-propanol. Conversions<br />

and ee’s were determined by GC (50 m Chiraldex �-PM,<br />

130 °C, (S)-7a 15.1 min, (R)-7a 16.4 min).<br />

Figure 3. Dependency of enantioselectivity versus E/Z ratio.<br />

Reactions were carried out at 10 °C for 24 h with 0.0024 mmol<br />

[Rh(cod)2]BF4, 0.005 mmol ligand 6a and 0.24 mmol substrate<br />

in 2.0 mL of ethanol. Conversions and ee’s were determined<br />

by GC (50 m Chiraldex �-PM, 130 °C, (S)-7a 15.1 min, (R)-7a<br />

16.4 min).<br />

crystallization is feasible, the selective hydrogenation of the<br />

mixture is clearly advantageous. This prompted us to explore<br />

the hydrogenation of different E/Z ratios including the<br />

mixture obtained from the synthesis (Figure 3). The results<br />

showed a linear decrease of enantioselectivity for all mixtures<br />

and demostrated that the hydrogenation of the single isomers<br />

led to highest enantioselectivity.<br />

Recently, Heller and Börner have reported kinetics and<br />

mechanistic investigations for the hydrogenation of E-7 and<br />

Z-7 through the use of bidentate phosphine ligands. The<br />

mentioned results modified the existent declaration that the


Figure 4. Dependency of hydrogen consumption versus reaction<br />

time. Reactions were carried out under isobaric conditions<br />

(1.0 bar) at 25 °C. [Rh(cod)2]BF4 (0.0072 mmol) and 0.015 mmol<br />

ligand 6a were stirred for 10 min under hydrogen atmosphere<br />

in 10.0 mL of methanol. Afterwards the substrate (0.72 mmol)<br />

was added in 1.0 mL of methanol. The conversions and ee’s<br />

were determined by GC (50 m Chiraldex �-PM, 130 °C, (S)-7a<br />

15.1 min, (R)-7a 16.4 min).<br />

reaction rate for hydrogenation of E-isomers is higher than<br />

that for the Z-isomers, because in some cases the opposite<br />

behaviour was observed. 3g<br />

To classify ligand 6a we recorded the hydrogen uptake<br />

in relation to the reaction time for the asymmetric hydrogenation<br />

of E-7 and Z-7 with our catalyst in isobaric conditions<br />

and under normal pressure of hydrogen (Figure 4).<br />

To minimize the interfering effect of cyclooctadiene (cod),<br />

the precatalyst was stirred for 10 min under hydrogen<br />

atmosphere before adding the substrate E-7 or Z-7. We were<br />

surprised to see that, unlike the case of most bidentate<br />

ligands, in our case at 50% conversion it was the Z-isomer<br />

which was hydrogenated faster (about 3.5 times the Eisomer).<br />

For gaining an insight into the structure of the catalyst,<br />

the hydrogenation of Z-7 was perfomed using the preformed<br />

complex [Rh(6a)2(nbd)] + CF3SO3 - as the catalyst. This<br />

cationic complex was prepared as previously reported by<br />

us, 11f and the reaction was run in MeOH at 5 bar at 25 °C.<br />

Under these conditions, Z-7 was completely hydrogenated<br />

in 12 h to give the (S)-enantiomer in 92% ee. This result is<br />

quite close to the one obtained with the catalysts prepared<br />

in situ by adding 2 equiv of the ligand 6a either to<br />

[Rh(nbd)2] + BF4 - (89% ee) or to [Rh(cod)2] + BF4 - (88% ee).<br />

From this it follows that the catalytically active species most<br />

likely contain two monodentate P-ligands per Rh center and<br />

that the ancillary diolefin ligand has a negligible effect, if<br />

any, on the stereoselectivity, while the anion may exert some<br />

influence. Further support to the presence of two ligands<br />

around the metal comes from the search of nonlinear effect<br />

(NLE) in the model reaction. 25 A set of hydrogenations was<br />

performed on both the isomers using various samples of<br />

Figure 5. Dependency of the optical purity of ligand 6a on<br />

product selectivity. Reactions were carried out at 10 °C for 24 h<br />

with 0.0024 mmol [Rh(cod)2]BF4, 0.005 mmol ligand 6a and<br />

0.24 mmol substrate in 2.0 mL of methanol. Conversions and<br />

ee’s were determined by GC (50 m Chiraldex �-PM, 130 °C,<br />

(S)-7a 15.1 min, (R)-7a 16.4 min).<br />

ligand 6a of different enantiomeric purity, and the ee’s<br />

observed have been plotted against the enantiomeric purity<br />

of the ligand (Figure 5). For both Z-7 and E-7 a positive<br />

nonlinear effect, which implies the bonding of at least two<br />

ligands to the rhodium, is clearly apparent. Similar results<br />

were reported by Reetz, 7i Zhou, 9b and Feringa 6f for the<br />

asymmetric hydrogenation of itaconic esters or R-amino acid<br />

derivatives by the use of different monodentate P-donor<br />

ligands. To verify the positive nonlinear effect we decreased<br />

the amount of ligand to 1 equiv with respect to rhodium.<br />

Here, we observed a diminished reaction rate compared to<br />

the rate of hydrogenation with 2 equiv of ligand. In addition,<br />

the amount of ligand was increased to 4 equiv with respect<br />

to rhodium, but also in this case a reduced reaction rate was<br />

monitored.<br />

In previous studies we have shown the pronounced effect<br />

that the substitution pattern at the P-centre of BINEPINE<br />

ligands has on the stereoselectivity of the asymmetric<br />

process. 13 In order to evaluate this substituent effect the<br />

asymmetric hydrogenation of E-7 and Z-7 was performed<br />

in the optimized conditions previously devised with nine<br />

different 4,5-dihydro-3H-dinaphtho[2,1-c;1′,2′-e]phosphepines<br />

(6a-6i) (E-7: 2.5 bar hydrogen pressure, 2-propanol,<br />

10 °C, 24 h; Z-7: 50.0 bar, ethanol, 10 °C, 24 h) (Table 1).<br />

In the hydrogenation of E-7 the best enantioselectivities<br />

up to 90, 88, and 87% ee, respectively, were achieved with<br />

ligands 6g, 6d, and much to our surprise, with 6i (Table 1,<br />

entries 7, 4, and 9). The last one has always given quite poor<br />

results in the other benchmark tests where it has been<br />

screened. The presence of electron-donating groups on the<br />

P-aryl substituent has a positive effect on the stereoselectivity<br />

(Table 1, entries 4, 6, and 7), whereas the opposite occurs<br />

for electron-withdrawing groups (Table 1, entry 2). This trend<br />

(25) (a) Girard, C.; Kagan, H. B.; Angew. Chem. 1998, 110, 3088-3127. (b)<br />

Faller, J. W.; Parr, J. J. Am. Chem. Soc. 1993, 115, 804-805.<br />

Vol. 11, No. 3, 2007 / Organic Process Research & Development • 573


Table 1. Hydrogenation of E-7 and Z-7 with different<br />

4,5-dihydro-3H-dinaphtho[2,1-c;1′,2′-e]phosphepines (6a-6i)<br />

entry ligand isomer a conv. [%] b ee [%] b isomer c conv. [%] b ee [%] b<br />

1 6a E >99 79 (R) Z >99 92 (S)<br />

2 6b E 91 40 (R) Z >99 32 (S)<br />

3 6c E 80 69 (R) Z >99 80 (S)<br />

4 6d E >99 88 (R) Z >99 86 (S)<br />

5 6e E >99 20 (R) Z >99 rac<br />

6 6f E 91 81 (R) Z 90 40 (S)<br />

7 6g E >99 90 (R) Z >99 89 (S)<br />

8 6h E >99 76 (R) Z 5 38(S)<br />

9 6i E >99 87 (R) Z 3 60(S)<br />

a All reactions were carried out at 10 °C under 2.5 bar pressure of hydrogen<br />

for 24 h with 0.0024 mmol [Rh(cod)2]BF4, 0.005 mmol ligand and 0.24 mmol<br />

substrate in 2-propanol (2.0 mL). b Conversions and ee’s were determined by<br />

GC (50 m Chiraldex �-PM, 130 °C, (S)-7a 15.1 min, (R)-7a 16.4 min). The<br />

absolute configuration was determined by comparing with reported data. 3l c All<br />

reactions were carried out at 10 °C under 50.0 bar pressure of hydrogen for 24<br />

h with 0.0024 mmol [Rh(cod)2]BF4, 0.005 mmol ligand and 0.24 mmol substrate<br />

in ethanol (2.0 mL).<br />

is similar to the one observed in the hydrogenation of Z-7<br />

although in that case the best ee was scored with the plain<br />

phenyl ligand 6a. Alkyl-substituted phosphepines 6h and 6i<br />

produced catalysts of negligible activity in hydrogenation<br />

of this substrate (Table 1, entries 8 and 9).<br />

To explore scope and limitation of our ligand toolbox,<br />

the asymmetric hydrogenation of a range of �-dehydroamino<br />

acids derivatives was performed under optimized conditions<br />

(Table 2). First the influence of the ester-group on the<br />

hydrogenation of both Z- and E-isomers was investigated in<br />

detail. For E-7 changing from methyl to isopropyl ester<br />

resulted consistently in a slight decrease of stereoselectivity,<br />

whereas in the case of the Z-isomer no reliable correlation<br />

between ester functionality and enantioselectivity was detected.<br />

Notably, a diminished yield for all hydrogenations<br />

of Z-9 was attained (Table 2).<br />

Finally, the influence of � 3 - and � 2 -substitution was<br />

investigated on a set of seven different substrates (Scheme<br />

4 and Table 3). While substitution of a methyl with an ethyl<br />

group has negligible effects, an increase in the branching of<br />

the alkyl group resulted in an improved selectivity with<br />

ligands 6g and 6i. Furthermore, we also carried out a<br />

substitution in the � 3 -position by an aromatic group (compound<br />

Z-12). As a tendency, depletion of conversion and<br />

enantioselectivity was observed in comparison to Z-8 (Table<br />

1, entries 1-7 and Table 2, entries 15-21). A similar<br />

negative effect was found after substitution of the acyl<br />

protecting group by benzoyl and subjecting Z-13 in the<br />

hydrogenation reaction (Table 3, entries 15-21).<br />

574 • Vol. 11, No. 3, 2007 / Organic Process Research & Development<br />

Table 2. Influence of ester group on conversion and<br />

enantioselectivity<br />

entry ligand isomer a conv. [%] b ee [%] b isomer c conv. [%] b ee [%] b<br />

1 6a E-8 98 83 (R) Z-8 >99 94 (S)<br />

2 6b E-8 79 46 (R) Z-8 >99 rac<br />

3 6c E-8 65 20 (R) Z-8 >99 36 (S)<br />

4 6d E-8 98 84 (R) Z-8 >99 83 (S)<br />

5 6g E-8 97 84 (R) Z-8 >99 85 (S)<br />

6 6h E-8 >99 66 (R) Z-8 >99 rac<br />

7 6i E-8 >99 80 (R) Z-8 >99 rac<br />

8 6a E-9 >99 70 (R) Z-9 52 78 (S)<br />

9 6b E-9 63 40 (R) Z-9 23 46 (S)<br />

10 6c E-9 73 20 (R) Z-9 61 24 (S)<br />

11 6d E-9 >99 76 (R) Z-9 69 91 (S)<br />

12 6g E-9 92 72 (R) Z-9 55 73 (S)<br />

13 6h E-9 >99 66 (R) Z-9 34 rac<br />

14 6i E-9 >99 76 (R) Z-9 32 6 (S)<br />

a All reactions were carried out at 10 °C under 2.5 bar pressure of hydrogen<br />

for 24 h with 0.0024 mmol [Rh(cod)2]BF4, 0.005 mmol ligand and 0.24 mmol<br />

substrate in 2-propanol (2.0 mL). b Conversions and ee’s were determined by<br />

GC (8 (25 m Lipodex E, 70/40-8-180, (R)-8a 33.2 min, (S)-8a 33.4 min), 9 (25<br />

m Lipodex E, 70/25-10-180, (R)-9a 33.2 min, (S)-9a 33.4 min). The absolute<br />

configurations were determined by comparing the sign of specific rotation with<br />

reported data. 3d c All reactions were carried out at 10 °C under 50.0 bar pressure<br />

of hydrogen for 24 h with 0.0024 mmol [Rh(cod)2]BF4, 0.005 mmol ligand and<br />

0.24 mmol substrate in ethanol (2.0 mL).<br />

Scheme 4. Asymmetric hydrogenation of methyl<br />

2-acetamidocylopent-1-enecarboxylate (14) a<br />

a All reactions were carried out at 10 °C under 50.0 bar pressure of hydrogen<br />

for 24 h with 0.0024 mmol [Rh(cod)2]BF4, 0.005 mmol ligand and 0.24 mmol<br />

substrate in ethanol (2.0 mL). Conversion, de’s and ee’s were determined by<br />

GC (25 m Chirasil Val, 120 °C, 14a1 26.8 min, 14a2 27.6 min, 14a3 52.1 min,<br />

14a4 53.0 min).<br />

As an example for tetra-substituted �-dehydroamino acid<br />

precursor (� 3 - and � 2 -substitution) we tested methyl2acetamidocyclopent-1-enecarboxylate<br />

(14) 26 under optimized<br />

conditions for Z-isomers (Scheme 4). In the majority of cases<br />

excellent diastereoselectivities up to >99% de were achieved,<br />

accompanied by good to moderate conversions. Best enantioselectivities<br />

up to 76% ee were obtained by utilizing<br />

ligands 6a and 6b (Scheme 4).<br />

An additional possibility offered by monodentate ligands<br />

from which one can profit in asymmetric catalysis is the<br />

possibility to introduce in the coordination sphere of the<br />

(26) Tang, W.; Wu, S.; Zhang, X. J. Am. Chem. Soc. 2003, 125, 9570-9571.


Table 3. Influence of � 3 -substitution on conversion and<br />

enantioselectivity<br />

entry ligand isomer a conv. [%] b ee [%] b isomer c conv. [%] b ee [%] b<br />

1 6a E-10 >99 78 (R) Z-10 98 86 (S)<br />

2 6b E-10 91 44 (R) Z-10 96 58 (S)<br />

3 6c E-10 95 60 (R) Z-10 >99 20 (S)<br />

4 6d E-10 >99 78 (R) Z-10 97 50 (S)<br />

5 6g E-10 >99 75 (R) Z-10 93 48 (S)<br />

6 6h E-10 >99 72 (R) Z-10 97 rac<br />

7 6i E-10 >99 84 (R) Z-10 47 24 (R)<br />

8 6a E-11 >99 26 (R) Z-11 >99 78 (S)<br />

9 6b E-11 39 42 (R) Z-11 99 68 (S)<br />

10 6c E-11 88 70 (R) Z-11 99 28 (S)<br />

11 6d E-11 50 60 (R) Z-11 94 26 (S)<br />

12 6g E-11 >99 88 (R) Z-11 94 68 (S)<br />

13 6h E-11 98 34 (R) Z-11 67 rac<br />

14 6i E-11 >99 90 (R) Z-11 10 58 (R)<br />

15 6a Z-12 99 70 (R) Z-13 19 70 (-)<br />

16 6b Z-12 46 26 (R) Z-13 9 4 (-)<br />

17 6c Z-12 90 47 (R) Z-13 7 46(+)<br />

18 6d Z-12 95 52 (R) Z-13 99 76 (S)<br />

3 6a PCy3 Z a >99 78 (S)<br />

4 6i 6i E b >99 87 (R)<br />

5 6i PPh3 E b >99 34 (R)<br />

6 6i PCy3 E b >99 88 (R)<br />

7 6i 15 d E b 21 12 (R)<br />

a All reactions were carried out at 10 °C under 50.0 bar pressure of hydrogen<br />

for 24 h with 0.0024 mmol [Rh(cod)2]BF4, 0.005 mmol ligand and 0.24 mmol<br />

substrate in ethanol (2.0 mL). b All reactions were carried out at 10 °C under<br />

2.5 bar pressure of hydrogen for 24 h with 0.0024 mmol [Rh(cod)2]BF4, 0.005<br />

mmol ligand and 0.24 mmol substrate in 2-propanol (2.0 mL). c Conversions<br />

and ee’s were determined by GC (50 m Chiraldex �-PM, 130 °C, (S)-7a 15.1<br />

min, (R)-7a 16.4 min). The absolute configuration was determined by comparing<br />

with reported data. 3l d 15 ) tris(2,4-di-tert-butylphenyl)phosphite.<br />

of 1.05 to 1.05 equiv of the two ligands with respect to 1.0<br />

equiv of [Rh(cod)2]BF4, and the reactions were carried out<br />

under the optimized conditions defined above. As a general<br />

trend, the heterocombination of monodentate ligands was<br />

always less stereoselective compared to the homocombination<br />

of the pivotal ligands (Table 4). There was just one<br />

notable exception for the system 6i/tricyclohexylphosphine<br />

(Table 4, entry 6) which gave the same ee as obtained when<br />

6i was used alone. Even if these results do not permit to<br />

draw any substantial conclusion, we may infer that in these<br />

conditions the catalytic activity of the different complexes<br />

which are formed in solution lies in the same range and that<br />

they compete for the substrate.<br />

Summary<br />

In conclusion, we have shown that monodentate chiral<br />

4,5-dihydro-3H-dinaphtho[2,1-c;1′,2′-e]phosphepine ligands<br />

can be synthesized on multi-10-g scale from 2,2′-binaphthol.<br />

They can be tuned easily by variation of the substituent at<br />

the P-atom. This class of ligands can be used for various<br />

rhodium- and ruthenium-catalyzed hydrogenations. For the<br />

first time their application towards the synthesis of �-amino<br />

acid derivatives is shown, and enantioselectivities up to 94%<br />

ee were achieved.<br />

Experimental Section<br />

All manipulations were performed under argon<br />

atmosphere using standard Schlenk techniques. Toluene,<br />

n-hexane, and diethyl ether were distilled from sodium<br />

benzophenone ketyl under argon. Methanol was distilled<br />

from magnesium under argon. Ethanol and 2-propanol were<br />

Vol. 11, No. 3, 2007 / Organic Process Research & Development • 575


distilled from sodium under argon. Methylene chloride was<br />

distilled from CaH2 under argon. Ligands 6 were synthesized<br />

according to our previously published protocols. 13 [Rh(cod)2]-<br />

BF4 (purchased from Fluka) was used without further<br />

purification. The synthesis of (S)-2,2′-dimethyl-1,1′-binaphthyl<br />

in a 200-g scale has been described in the literature. 12b<br />

Synthesis of 4-Phenyl-4,5-dihydro-3H-dinaphtho[2,1c;1′,2′-e]phosphepine<br />

(6a) and 4-tert-Butyl-4,5-dihydro-<br />

3H-dinaphtho[2,1-c;1′,2′-e]phosphepine (6i) on 10-g Scale.<br />

Synthesis of the Dilithio Species. A solution of n-BuLi (0.19<br />

mol, 120 mL, 1.6 M in n-hexane) was concentrated under<br />

vacuum and the residual oil dissolved in diethyl ether (60<br />

mL). After cooling the mixture to 0 °C a solution of (S)-<br />

2,2′-dimethyl-1,1′-binaphthyl (74.5 mmol, 21 g) in diethyl<br />

ether (140 mL) was added over a dropping funnel during 20<br />

min to give a red solution. Afterwards TMEDA (192 mmol,<br />

28.6 mL, distilled over CaH2) was added slowly, and the<br />

resulting solution was kept for 24 h at room temperature,<br />

yielding deep red crystals. The supernatant solution was<br />

decanted via a tube. The crystals were washed two times<br />

with dry n-hexane (30 mL, removed by a tube) and dried<br />

under vacuum to give 60-80% yield (24.6-37.8 g).<br />

Synthesis of the 4-Phenyl-4,5-dihydro-3H-dinaphtho[2,1c;1′,2′-e]phosphepine<br />

(6a). Starting with the dilithium salt of<br />

(S)-2,2′-dimethyl-1,1′-binaphthyl (24.5 g, 44.4 mmol) in n-hexane<br />

(130 mL) a solution of phenyl dichlorophosphine (6.8 mL,<br />

50.3 mmol) in n-hexane (55 mL) was added at 0 °C. After<br />

2 h refluxing, the reaction mixture was quenched with water/<br />

toluene. The organic layer was separated and dried over<br />

MgSO4. The ligand 6a was purified by column chromatography<br />

in dry toluene and gave light-yellow foam (12.9 g; 75%).<br />

Synthesis of the 4-tert-Butyl-4,5-dihydro-3H-dinaphtho-<br />

[2,1-c;1′,2′-e]phosphepine (6i). Starting with the dilithium<br />

salt of (S)-2,2′-dimethyl-1,1′-binaphthyl (24.5 g, 44.4 mmol)<br />

in n-hexane (130 mL) a solution of tert-butyl dichlorophosphine<br />

(8 g, 50.3 mmol) in n-hexane (55 mL) was added at<br />

0 °C. After 3 h refluxing, the reaction mixture was quenched<br />

with water/toluene. The organic layer was separated and dried<br />

over MgSO4. The ligand 6i was purified by crystallization<br />

from toluene (13.2 g; 81%).<br />

General Synthesis of �-Dehydroamino Acid Derivatives<br />

7-11 and 14. 3c To a solution of NH4OAc (0.36 mol)<br />

in methanol, ethanol, or 2-propanol (depending on the esterfunctionality<br />

(100 mL)) was added the corresponding �-ketoester<br />

(0.07 mol). After stirring for 60 h at room temperature<br />

the solvent was removed, and a mixture of CHCl3 and water<br />

was added. The organic layer was washed with water (3 ×<br />

50 mL) and brine (50 mL) and dried over Na2SO4. The<br />

solvent was evaporated under reduced pressure, yielding the<br />

corresponding 3-amino-2-alkenoate. The 3-amino-2-alkenoate<br />

(0.07 mol), pyridine (0.13 mol), and acetic anhydride<br />

(0.32 mol) were dissolved in THF (50 mL) and stirred for<br />

12hat70°C (procedure A) or 95 °C (procedure B). The<br />

solution was reduced to half of the volume and ethylacetate<br />

was added. After washing with water, HCl, NaHCO3, brine<br />

and drying over Na2SO4, the solvent was removed.<br />

Procedure A (E-Isomer Enriched Residue). Dissolving the<br />

residue in ethylacetate/n-hexane (1:1) and storing the solution<br />

576 • Vol. 11, No. 3, 2007 / Organic Process Research & Development<br />

overnight at -20 °C yielded the E-isomers of �-dehydroamino<br />

acid derivatives 7-11, which were recrystallized<br />

three times from ethylacetate/n-hexane (1:1) to give colorless<br />

crystals [yield of the main fraction referred to �-ketoester:<br />

E-7: 1.4 g (11%), E-8: 1.2 g (10%), E-9: 1.5 g (11%),<br />

E-10: 1.4 g (12%), E-11: 1.8 g (14%)].<br />

Procedure B (Z-isomer Enriched Residue). The pure<br />

Z-isomer was obtained by column chromatography (eluent:<br />

ethylacetate/n-hexane, 1:1 or 1:2) [yield referred to �-ketoester:<br />

Z-7: 4.3 g (39%), Z-8: 1.2 g (35%) (0.035 mmol<br />

�-ketoester), Z-9: 6.0 g (46%), Z-10: 1.6 g (26%) (0.035<br />

mmol �-ketoester), Z-11: 4.0 g (31%)]. In the case of methyl<br />

2-acetamidocyclopent-1-enecarboxylate (14) purification was<br />

carried out by crystallization (3×) from ethylacetate/n-hexane<br />

(1:1), yielding brown crystals [yield of the main fraction:<br />

4.1 g (32%)].<br />

Synthesis of Ethyl 3-Acetamido-3-phenyl-2-propenoate<br />

(Z-12). 3l To a solution of NH4OAc (0.65 mol) in ethanol<br />

(100 mL) was added the corresponding �-ketoester (0.13<br />

mol). After stirring for 60 h at room temperature the solvent<br />

was removed, and a mixture of CHCl3 and water was added.<br />

The organic layer was washed with water (3 × 50 mL) and<br />

brine (50 mL) and dried over Na2SO4. The solvent was<br />

evaporated under reduced pressure, yielding the corresponding<br />

3-amino-2-alkenoate. To a stirred solution of the corresponding<br />

3-amino-2-alkenoate (0.026 mmol) and pyridine<br />

(0.046 mol) in toluene (50 mL) was added dropwise a<br />

solution of acetyl chloride (0.027 mol) in toluene (10 mL)<br />

at 0 °C. The mixture was stirred for 48 h at 25 °C. A further<br />

amount of acetylchloride (0.027 mol) in toluene (10 mL)<br />

was added at 0 °C, and the resulting mixture was stirred for<br />

48 h. After quenching with aqueous NH4H2PO4 solution the<br />

mixture was extracted with ethylacetate (3 × 50 mL). The<br />

organic layer was washed with water (2 × 50 mL) and dried<br />

over Na2SO4. The solvents were removed in vacuo, and the<br />

obtained yellow oil was purified by column chromatography<br />

(eluent: ethylacetate/n-hexane, 1:1), yielding yellow crystals<br />

[yield: Z-12: 0.59 g (10%)].<br />

Synthesis of Methyl 3-Benzamido-2-butenoate (Z-13). 3l<br />

To a solution of NH4OAc (0.65 mol) in ethanol (100 mL)<br />

was added the corresponding �-ketoester (0.13 mol). After<br />

stirring for 60 h at room temperature the solvent was<br />

removed, and a mixture of CHCl3 and water was added. The<br />

organic layer was washed with water (3 × 50 mL) and brine<br />

(50 mL) and was dried over Na2SO4. The solvent was<br />

evaporated under reduced pressure, yielding the corresponding<br />

3-amino-2-alkenoate. To a stirred solution of the corresponding<br />

3-amino-2-alkenoate (0.026 mmol) and pyridine<br />

(0.046 mol) in diethylether (70 mL) was added dropwise a<br />

solution of benzoyl chloride (0.027 mol) in toluene (10 mL)<br />

at -30 °C. The mixture was stirred for 12 h at -30 °C and<br />

for 24 h at room temperature. After quenching with aqueous<br />

water the mixture was extracted with ethylacetate (3 × 50<br />

mL). The organic layer was washed with HCl (1 mol/L) and<br />

brine and dried over Na2SO4. The solvents were removed in<br />

vacuo, and the obtained yellow oil was purified by column<br />

chromatography (eluent: ethylacetate/n-hexane, 1:1), yielding<br />

crystals [yield: Z-13: 2.9 g (9%)].


General Procedure for the Catalytic Hydrogenation<br />

of �-Dehydroamino Acid Derivatives. A solution of the<br />

�-dehydroamino acid derivative (0.24 mmol) and 1.0 mL of<br />

solvent was transferred via syringe into an autoclave charged<br />

with argon. The catalyst was generated in situ by stirring<br />

[Rh(cod)2]BF4 (0.0024 mmol) and the corresponding 4,5dihydro-3H-dinaphthophosphepine<br />

ligand (0.005 mmol) in<br />

1.0 mL of solvent for a period of 10 min and afterwards<br />

transferring via syringe into the autoclave. The autoclave was<br />

then charged with hydrogen and stirred at the required<br />

temperature. After the predetermined time the hydrogen was<br />

released, and the reaction mixture passed through a short<br />

plug of silica gel. The conversion was measured by GC or<br />

1 H NMR and the enantioselectivity by GC or HPLC.<br />

Acknowledgment<br />

We thank Mrs. M. Heyken, Mrs. C. Voss, Mrs. G.<br />

Wenzel, Mrs. K. Schröder, Mrs. S. Buchholz, and Dr. C.<br />

Fischer (all Leibniz Institut für Katalyse e.V. an der<br />

<strong>Universität</strong> <strong>Rostock</strong>) for excellent technical and analytical<br />

assistance. Dr. B. Hagemann is grate<strong>full</strong>y thanked for the<br />

syntheses of ligands 6e and (R)-6a. Generous financial<br />

support from the state of Mecklenburg-Western Pomerania<br />

and the BMBF as well as the Deutsche Forschungsgemeinschaft<br />

(Leibniz-Price) is grate<strong>full</strong>y acknowledged.<br />

Received for review October 30, 2006.<br />

OP0602270<br />

Vol. 11, No. 3, 2007 / Organic Process Research & Development • 577


5. Application of monodentate phosphine ligands in asymmetric hydrogenation 87<br />

5.3. Enantioselective rhodium–catalyzed hydrogenation of enol<br />

carbamates in the presence of monodentate phosphines<br />

Stephan Enthaler, Giulia Erre, Kathrin Junge, Dirk Michalik, Anke Spannenberg,<br />

Serafino Gladiali, and Matthias Beller*, Tetrahedron: Asymmetry, 2007, 18, 1288–<br />

1298.<br />

Contributions<br />

The synthetic route to ligand 7 was carried by G. Erre via compound 5a and 6. Suitable x–ray<br />

crystals of compound 5b were grown by Dr. K. Junge, while crystals for oxide 6 were<br />

obtained by G. Erre. The synthesis of substrates 9, 10, 11, 14 and the corresponding<br />

hydrogenation experiments stated in table 2 and table 3 were performed by G. Erre.


Enantioselective rhodium-catalyzed hydrogenation of enol carbamates<br />

in the presence of monodentate phosphines<br />

Stephan Enthaler, a Giulia Erre, a Kathrin Junge, a Dirk Michalik, a Anke Spannenberg, a<br />

Fabrizio Marras, b Serafino Gladiali b and Matthias Beller a, *<br />

a Leibniz-Institut für Katalyse e.V. an der <strong>Universität</strong> <strong>Rostock</strong>, Albert-Einstein-Str. 29a, 18059 <strong>Rostock</strong>, Germany<br />

b Dipartimento di Chimica, Università di Sassari, Via Vienna 2, 07100 Sassari, Italy<br />

Received 27 March 2007; accepted 1 June 2007<br />

Abstract—The rhodium-catalyzed asymmetric hydrogenation of different acyclic and cyclic enol carbamates to give optically active carbamates<br />

has been examined in the presence of chiral monodentate ligands based on a 4,5-dihydro-3H-dinaphthophosphepine motif 4.<br />

The enantioselectivity is largely dependent upon the reaction conditions, the nature of substituents on the phosphorus ligand and structure<br />

of the enol carbamate. By applying the optimized reaction conditions, enantioselectivities of up to 96% ee have been achieved.<br />

Ó 2007 Elsevier Ltd. All rights reserved.<br />

1. Introduction<br />

The importance of enantiomerically pure compounds is<br />

demonstrated by their widespread application as either<br />

building blocks or intermediates for the synthesis of pharmaceuticals,<br />

agrochemicals, polymers, natural compounds,<br />

auxiliaries, ligands and synthons in organic syntheses. 1 To<br />

serve the growing demand for enantiomerically pure compounds,<br />

various synthetic approaches have been developed.<br />

Within the different molecular transformations to chiral<br />

compounds, catalytic reactions offer an efficient and<br />

versatile strategy and present a key technology for the<br />

advancement of ‘green chemistry’, specifically for waste<br />

prevention, reducing energy consumption, achieving high<br />

atom efficiency and generating advantageous economics. 2<br />

In this regard, the use of molecular hydrogen in reductions<br />

of C@C, C@O andC@N bonds is one of the most extensively<br />

studied fields. For activation of the hydrogen and<br />

transfer of chirality, the use of transition metal catalysts<br />

containing chiral ligands are essential. Over a period of<br />

nearly 30 years, the synthesis of new ligands focused mainly<br />

on bidentate phosphorus systems, because of the promising<br />

results in the first few years of homogeneous asymmetric<br />

hydrogenation. However, at the end of the 20th century,<br />

the situation changed and monodentate ligands received<br />

* Corresponding author. Tel.: +49 381 1281 113; fax: +49 381 1281<br />

5000; e-mail: Matthias.Beller@catalysis.de<br />

0957-4166/$ - see front matter Ó 2007 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.tetasy.2007.06.001<br />

Tetrahedron: Asymmetry 18 (2007) 1288–1298<br />

more attention. 3,4 The advantages of monodentate phosphorus<br />

ligands compared to bidentate ligands are their easier<br />

synthesis, high tunability and even a higher efficiency in<br />

the transfer of chiral information as observed in several<br />

cases. Important contributions in this field with excellent<br />

enantioselectivities for various classes of substrates were reported<br />

by Feringa and de Vries et al. 5 (phosphoramidites),<br />

Reetz et al. 6 (phosphites), Pringle and Claver et al. 7 (phosphonites)<br />

and Zhou et al. (spiro phosphoramidites). 8<br />

Following the effort of one of us (S.G.) 9 and parallel to the<br />

work of Zhang, 10 we constructed a ligand library of<br />

approximately 25 different monodentate phosphines based<br />

on a 4,5-dihydro-3H-dinaphtho[2,1-c;1 0 ,2 0 -e]phosphepine<br />

framework 4 (Scheme 1), which is reminiscent of the 1,1 0 -<br />

binaphthol core, which is present in the ligands mentioned<br />

above (Scheme 1). 11 The potential of this ligand class 4 in<br />

asymmetric hydrogenation with molecular hydrogen, such<br />

as the reduction of a-amino acid precursors, dimethyl<br />

itaconate, enamides and b-ketoesters achieving enantioselectivities<br />

up to 95% ee was previously reported by us. 11<br />

Moreover, other groups demonstrated the usefulness of<br />

these ligands in several catalytic asymmetric reactions. 12<br />

The promising results obtained in the asymmetric hydrogenation<br />

of N-(1-phenylvinyl)acetamides with enantioselectivities<br />

of up to 95% ee and catalyst activities up to<br />

2000 h 1 (TOF) motivated us to explore our ligand library<br />

4 in the hydrogenation of structurally similar enol carbamates,<br />

which offer an alternative approach to chiral


1. n-BuLi/<br />

TMEDA<br />

2. Cl 2PNEt 2<br />

HCl<br />

alcohols. Pioneering work in the field of enantioselective<br />

hydrogenation of enol carbamates has been reported by<br />

Feringa, de Vries and Minnaard et al. who have scored<br />

enantioselectivities of up to 98% ee with rhodium-catalysts<br />

containing monodentate phosphoramidites (MonoPhosfamily).<br />

2.1. Ligand synthesis<br />

P NEt 2<br />

2<br />

2. Results and discussion<br />

In general, the ligands were prepared by double metallation<br />

of 2,2 0 -dimethylbinaphthyl 1 with 2 equiv of n-butyl lithium,<br />

followed by quenching with diethylaminodichlorophosphine<br />

or with aryl and alkyl dichlorophosphines.<br />

Deprotection of the diethylamino phosphine 2 with gaseous<br />

HCl produced 4-chloro-4,5-dihydro-3H-dinaphtho[2,1c;1<br />

0 ,2 0 -e]phosphepine 3 in 80% yield. This enantiomerically<br />

pure chlorophosphine is easily coupled with various Grignard<br />

reagents or organo-lithium compounds to give a broad<br />

selection of ligands 4. In order to expand the application of<br />

this class of ligands, we prepared derivatives of 1, bearing<br />

CH 3<br />

CH 3<br />

P Cl<br />

3<br />

1<br />

RMgX<br />

or RLi<br />

1. n-BuLi/<br />

TMEDA<br />

2. Cl 2PR<br />

P R<br />

4<br />

4a: R = Ph<br />

4b: R = p-CF 3C 6H 4<br />

4c: R = 3,5-(CH 3) 2C 6H 3<br />

4d: R = p-CH 3OC 6H 4<br />

4e: R = o-CH 3OC 6H 4<br />

4f: R = iPr<br />

4g: R = tBu<br />

Scheme 1. General synthesis of ligands with 2,2 0 -binaphthyl framework.<br />

S. Enthaler et al. / Tetrahedron: Asymmetry 18 (2007) 1288–1298 1289<br />

H 2O 2<br />

P Ph<br />

O<br />

P R<br />

7<br />

HSiCl 3<br />

6'<br />

5' 4'<br />

10' 3'<br />

7<br />

6<br />

5<br />

6<br />

10 2'<br />

12'<br />

7'<br />

8'<br />

8<br />

9'<br />

9<br />

1' 11'<br />

i<br />

5 11<br />

2<br />

4<br />

3<br />

12<br />

o O<br />

P m<br />

p<br />

LDA/MeI<br />

5a: R = Ph<br />

5b: R = p-CH 3OC 6H 4<br />

substituents on the aliphatic carbon at the a-position to<br />

the phosphorus. The desired bis-methylated product 7 has<br />

been obtained via a deprotonation–alkylation protocol.<br />

After the preliminary step of oxidation of 4-phenyl-4,5dihydro-3H-dinaphtho[2,1-c;1<br />

0 ,2 0 -e]phosphepine 4a with<br />

H2O2, lithiation and alkylation were performed by the addition<br />

of LDA (3 equiv) and CH 3I (2 equiv) at room temperature.<br />

The reaction is completely stereoselective, and only<br />

one of the possible dialkylated products is obtained<br />

(Scheme 1).<br />

Recently, Widhalm et al. reported a similar approach to<br />

the synthesis of compound 7, via the phosphepine sulfide<br />

instead of the oxide. 13 To compare the two procedures,<br />

the stereochemistry of the new stereogenic centres in the<br />

phosphepine oxide 6 was studied in detail.<br />

The stereochemistry of compound 6 was confirmed by<br />

NMR spectroscopy. The initial 31 P NMR measurements<br />

displayed the occurrence of exclusively one enantiomer<br />

for 6 and 7 in each case, since only one single signal was<br />

found and racemization of the binaphthyl backbone over<br />

the course of reactions was excluded. In a two-dimensional<br />

Table 1. Selected bond lengths [A˚ ] and angles [°] of the phosphepine oxides 5a, 5b and 6<br />

5a 6 5b a<br />

P1–C1 1.812(5) P1–C25 1.799(3) P1–C1 1.824(4)<br />

[P2–C30] [1.825(4)]<br />

P1–C7 1.813(4) P1–C22 1.832(2) P1–C29 1.833(4)<br />

[P2–C58] [1.846(4)]<br />

P1–C28 1.814(5) P1–C1 1.834(2) P1–C8 1.837(4)<br />

[P2–C37] [1.819(4)]<br />

P1–O1 1.481(3) P1–O1 1.483(2) P1–O1 1.444(3)<br />

[P2–O3] [1.407(5)]<br />

C7–P1–C1 109.6(2) C25–P1–C22 108.6(1) C1–P1–C29 107.9 (2)<br />

[C30–P2–C37] [107.2 (2)]<br />

C28–P1–C1 104.0(2) C25–P1–C1 104.2(1) C1–P1–C8 105.4(2)<br />

[C30–P2–C58] [103.0(2)]<br />

C7–P1–C28 103.1(2) C22–P1–C1 108.8 (1) C29–P1–C8 102.3(2)<br />

[C58–P2–C37] [100.8(2)]<br />

a<br />

Values of the second molecule in the asymmetric unit are in brackets.


1290 S. Enthaler et al. / Tetrahedron: Asymmetry 18 (2007) 1288–1298<br />

Figure 1. ORTEP plot of compounds 5a, 6 and 5b. Hydrogen atoms are omitted for clarity. The thermal ellipsoids correspond to 30% probability. With<br />

respect to compound 5b, only one of the two symmetry-independent molecules of the asymmetric unit is depicted.<br />

NOESY spectrum, correlations between the following protons<br />

were observed: o-Ph with H-3 0 , H-11 0 and Me(12); H-3<br />

with H-11 and H-3 0 with H-11 0 , respectively, indicating the<br />

bis-axial orientation of both methyl groups. The spatial<br />

proximity of the methyl group Me(12) and of the phenyl<br />

group is also reflected by the difference in 1 H chemical shift<br />

values of both methyl groups (dMe(12) = 0.61 and<br />

Me(12 0 ) = 0.95). The high-field shift of the signal for the<br />

methyl group Me(12) can be explained by the anisotropic<br />

effect of the phenyl substituent on the phosphorus.<br />

In addition, useful structural data were gained by X-ray<br />

crystallography of oxide 6. For comparison, X-ray structure<br />

analyses of the corresponding unsubstituted systems<br />

5a and 5b were determined. Selected bond lengths and angles<br />

are shown in Table 1. The molecular structures of the<br />

three oxides are given in Figure 1. In the case of 5b, two<br />

molecules have been found in the asymmetric unit with<br />

similar structural features. As expected in all compounds<br />

the substituents show an approximately tetrahedral<br />

arrangement at the phosphorus atom. The a,a-disubstituted<br />

phosphepine oxide 6 features shorter exocyclic<br />

(P1–C25 1.799(3) A ˚ ) than the endocyclic P–C bonds<br />

(P1–C22 1.832(2) A ˚ and P1–C1 1.834(2) A ˚ ) in comparison<br />

with the unsubstituted system 5a where all P–C bonds are<br />

equally long (Table 1). Furthermore, the a,a-substitution<br />

promotes a pronounced expansion of the endocyclic<br />

C–P–C angle up to 108.8(1)° (6) from 103.1(2)° (5a).<br />

In accordance with the NMR studies and the X-ray structure<br />

analysis, the configuration of compound 6 was assigned<br />

to be (S,S,Sa). The stereochemistry of ligand 7 was allotted<br />

in analogy to the one of the corresponding oxide.<br />

2.2. Catalytic experiments<br />

The synthesis of enol carbamates 8–14 was carried out in<br />

analogy to literature protocols, by reacting the corresponding<br />

ketone with NaH and subsequent addition of the corresponding<br />

carbamoyl chloride. 5m,14<br />

Initial studies on the influence of reaction conditions were<br />

performed with 1-phenylvinyl N,N-diethylcarbamate 8<br />

as substrate and 4-phenyl-4,5-dihydro-3H-dinaphtho[2,1c;1<br />

0 ,2 0 -e]phosphepine 4a as our standard ligand. Typically,<br />

we used an in situ pre-catalytic mixture of 1.0 mol %<br />

[Rh(cod)2]BF4 and 2.1 mol % of the ligand. All hydrogenation<br />

reactions were carried out in an eightfold parallel reactor<br />

array with a 3.0 ml reactor volume. 15<br />

First we investigated the influence of different solvents,<br />

such as methylene chloride, methanol, ethanol and 2-propanol<br />

in combination with a variation of the initial hydrogen<br />

pressure (1.0, 5.0, 10.0, 25.0 and 50.0 bar). Selected<br />

results are presented in Figure 2.<br />

ee [90%]<br />

100<br />

O<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

N<br />

O<br />

8<br />

1<br />

5<br />

[Rh(cod) 2]BF 4 + 2.1 (4a)<br />

25 °C, 24 h, H 2<br />

10<br />

pressure [bar]<br />

O O<br />

Surprisingly, by applying toluene as solvent in the hydrogenation<br />

of enol carbamate 8 no reaction occurred, while<br />

in previous studies toluene was found to be the solvent of<br />

choice for the hydrogenation of enamides and a-amino<br />

acids. 11c,e The best enantioselectivities of up to 96% ee were<br />

obtained with methanol as solvent with a hydrogen pressure<br />

of 25 bar.<br />

Two different pressure-enantioselectivity-dependencies<br />

have been noticed. In ethanol and 2-propanol, the enantioselectivity<br />

decreased when increasing the hydrogen pressure<br />

(1 bar: ethanol: 74% ee, 2-propanol: 74% ee, 50 bar: ethanol:<br />

25<br />

50<br />

N<br />

CH2Cl2 2-PrOH<br />

8a<br />

MeOH<br />

EtOH<br />

Figure 2. Solvent and pressure variation. Reactions were carried out at<br />

25 °C for 24 h with 0.0024 mmol [Rh(cod) 2]BF 4, 0.005 mmol ligand 4a and<br />

0.24 mmol 8 in 2.0 ml solvent.


64% ee, 2-propanol: 50% ee) whereas in methylene chloride<br />

and methanol, the reverse behaviour was observed (1 bar:<br />

methanol: 72% ee, methylene chloride: 66% ee, 50 bar:<br />

methanol: 92% ee, methylene chloride: 86% ee). With the<br />

exception of 2-propanol (>99%), moderate to good conversions<br />

(60–90%) were observed at lower pressures within a<br />

reasonable time (24 h), while at higher pressure a complete<br />

conversion was achieved. The results indicated methanol as<br />

the solvent of choice (conversion: >99%; enantioselectivity:<br />

96% ee). In addition, in order to estimate the effect of the<br />

protic solvent on the product structure the reaction was<br />

performed in methanol-d4. Analysis of the 1 H NMR spectra<br />

showed that no incorporation of deuterium in the product<br />

had occurred. Therefore, one can rule out any<br />

interference of the solvent in the catalytic cycle due to<br />

transfer hydrogenation and/or protonolysis of Rh-Cspecies.<br />

16<br />

In order to assess the optimum reaction conditions, a<br />

detailed pressure investigation was performed (Fig. 3).<br />

The results illustrated a constant range of enantioselectivity<br />

( 74% ee) between initial pressures of 1–18 bar, followed<br />

by an increase of up to a maximum of 96% ee at 25 bar.<br />

A further increase in the hydrogen pressure did not result<br />

in any improvement while enantioselectivities ranged<br />

between 92% and 94% ee in the region of 30–50 bar.<br />

ee [%]<br />

O<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

N<br />

O<br />

8<br />

[Rh(cod) 2]BF 4 + 2.1 (4a)<br />

25 ºC, 24 h, H 2<br />

O O<br />

1 5 10 12 15 18 20 25 30 40 50<br />

pressure [bar]<br />

Figure 3. Study of pressure-enantioselectivity-dependency. Reactions were<br />

carried out at 25 °C for 24 h with 0.0024 mmol [Rh(cod)2]BF4, 0.005 mmol<br />

ligand 4a and 0.24 mmol 8 in 2.0 ml methanol.<br />

The optimal reaction conditions devised in the solventpressure<br />

investigation (methanol as solvent and 25 bar as<br />

initial hydrogen pressure) were applied in a temperature<br />

study (Fig. 4). With the model ligand 4a in the hydrogenation<br />

of compound 8, we observed a high stability of the<br />

enantioselectivity (94–96% ee) in the range of 10–90 °C<br />

albeit with a reduction of conversion at 90 °C. A further<br />

increase to 110 °C produced a racemic mixture of 8a, probably<br />

as a result of ligand degradation. 17<br />

In order to evaluate the substituent effect at the phosphorus<br />

atom of the ligand, we studied the model reaction of 1-phen-<br />

S. Enthaler et al. / Tetrahedron: Asymmetry 18 (2007) 1288–1298 1291<br />

N<br />

8a<br />

[%]<br />

O<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

N<br />

O<br />

8<br />

[Rh(cod) 2]BF 4 + 2.1 (4a)<br />

24 h, H 2 (25 bar)<br />

O O<br />

10 30 50 70 90 110<br />

temperature [˚C]<br />

ylvinyl N,N-diethylcarbamate 8 in the presence of eight<br />

different 4,5-dihydro-3H-dinaphtho[2,1-c;1 0 ,2 0 -e]phosphepines<br />

4a–4g and 7 using the optimized reaction conditions<br />

(2.5 bar hydrogen pressure, methanol, 30 °C, 6 h). The<br />

results are presented in Table 2. In the hydrogenation of<br />

8, the best enantioselectivities of up to 90% and 72% ee,<br />

respectively, were achieved with ligand 4a and 4d (Table<br />

2, entries 1 and 4). The presence of electron-donating<br />

groups, as well as electron-withdrawing groups onto the<br />

P-aryl substituent, led to a significant decrease of the<br />

stereoselectivity (Table 1, entries 2–5).<br />

Catalysts containing alkyl-substituted phosphepines 4f or<br />

4g gave a significant lower activity and enantioselectivity<br />

compared to aryl-substituted phosphepines. Interestingly,<br />

in the case of ligand 4g and 7, the opposite enantiomer is<br />

formed during the reaction. A similar behaviour of the<br />

N<br />

8a<br />

ee<br />

conversion<br />

Figure 4. Dependency of enantioselectivity versus temperature. Reactions<br />

were carried out at corresponding temperature for 1–24 h with<br />

0.0024 mmol [Rh(cod) 2]BF 4, 0.005 mmol ligand 4a and 0.24 mmol 8 in<br />

2.0 ml methanol.<br />

Table 2. Hydrogenation of compound 8 with different 4,5-dihydro-3Hdinaphtho[2,1-c;1<br />

0 ,2 0 -e]phosphepines 4a–4g and ligand 7 a<br />

O<br />

N<br />

O<br />

8<br />

[Rh(cod) 2]BF 4 + 2.1 (4 or 7)<br />

MeOH, 6 h, H 2 (25 bar)<br />

N<br />

O O<br />

Entry Ligand Conv. (%) Yield (%) ee (%)<br />

1 4a >99 >99 96 (S)<br />

2 4b 40 40 28 (S)<br />

3 4c >99 >99 66 (S)<br />

4 4d >99 >99 72 (S)<br />

5 4e >99 >99 51 (S)<br />

6 4f 40 40 12 (S)<br />

7 4g 41 41 50 (R)<br />

8 b<br />

7 >99 >99 42 (R)<br />

a All reactions were carried out at 25 °C under 25 bar pressure of hydrogen<br />

for 6 h with 0.0024 mmol [Rh(cod)2]BF4, 0.005 mmol ligand and<br />

0.24 mmol 8 in methanol (2.0 ml).<br />

b 24 h.<br />

8a


1292 S. Enthaler et al. / Tetrahedron: Asymmetry 18 (2007) 1288–1298<br />

tert-butyl ligand 4g was previously reported for other<br />

asymmetric hydrogenations. 11c,f<br />

To explore the scope and limitation of our ligand toolbox,<br />

the asymmetric hydrogenation of a range of enol carbamates<br />

was performed under optimized conditions (Tables<br />

3 and 4). First the influence of the substitution at the carbamoyl-nitrogen<br />

was investigated (Table 3).<br />

Table 3. Variation of the protecting group in the substrate moiety a<br />

R1 R2 N<br />

O<br />

O<br />

[Rh(cod) 2]BF 4 + 2.1 (4 and 7)<br />

9: R 1=R 2=Me<br />

10: R 1=Me R 2=Ph<br />

11: R 1=Ph R 2=Ph<br />

MeOH, 24 h, H 2 (25 bar)<br />

R1 R2 N<br />

O O<br />

9a-11a<br />

Entry Ligand R1 R2 Conv. (%) Yield (%) ee (%)<br />

1 4a Me Me >99 >99 68 (S)<br />

2 4b Me Me 33 33 26 (S)<br />

3 4c Me Me 87 87 65 (S)<br />

4 4d Me Me >99 >99 75 (S)<br />

5 4e Me Me 99 99 40 (S)<br />

6 4f Me Me 26 26 17 (S)<br />

7 4g Me Me 44 44 33 (R)<br />

8 7 Me Me 80 80 27 (R)<br />

9 4a Me Ph >99 >99 58 (S)<br />

10 4b Me Ph 98 98 37 (S)<br />

11 4c Me Ph >99 >99 61 (S)<br />

12 4d Me Ph >99 >99 65 (S)<br />

13 4e Me Ph >99 >99 42 (S)<br />

14 4f Me Ph 53 53 4 (S)<br />

15 4g Me Ph >99 >99 45 (R)<br />

16 7 Me Ph >99 >99 53 (R)<br />

17 4a Ph Ph >99 88 10 (S)<br />

18 4b Ph Ph 98 77 Rac<br />

19 b<br />

4c Ph Ph 89 75 41 (S)<br />

20 4d Ph Ph >99 87 11 (S)<br />

21 4e Ph Ph >99 87 55 (S)<br />

22 4f Ph Ph 98 62 4 (S)<br />

23 4g Ph Ph >99 98 70 (R)<br />

24 7 Ph Ph >99 98 76 (R)<br />

a All reactions were carried out at 25 °C under 25 bar pressure of hydrogen<br />

for 24 h with 0.0024 mmol [Rh(cod) 2]BF 4, 0.005 mmol ligand and<br />

0.24 mmol substrate in methanol (2.0 ml).<br />

b Reaction was carried out for 6 h.<br />

All substrates were hydrogenated under the optimized conditions<br />

with eight different ligands 4a–4g and 7. The enol<br />

carbamates 9–11 undergo the reaction with good conversion<br />

in most cases (Table 3). The negative influence on<br />

the activity was again observed by the electron-withdrawing<br />

substituent on the phenyl (entries 2, 10 and 18) and<br />

by the alkyl group on the phosphorus, mainly i-Pr (Table<br />

3, entries 6, 7, 14 and 22). On the other hand in some cases<br />

a positive effect on the selectivity by substitution with electron<br />

donating groups was observed (Table 3, entries 4 and<br />

12). Again the stereoselectivity was reversed in the case of<br />

ligand 7 and 4g, with 7 being the higher inducer of chirality<br />

for system 11. The overall results for the three substrates<br />

9–11 are comparable, showing little influence of the differ-<br />

ent substituents on the carbamoyl nitrogen on the reaction<br />

path.<br />

Finally, the sensitivity on the variations in the double bond<br />

was investigated. Table 4 summarizes the results for three<br />

substrates with different substitutions: substitution of the<br />

phenyl-group by an alkyl-group 14, substitution in the 2position<br />

of the 1,1-olefin 12 and a cyclic substrate 13. Substrate<br />

Z-12 was hydrogenated with excellent conversion<br />

and good enantioselectivity by ligand 4d, while 13, which<br />

is structurally related to the E-isomer of substrate 12, fails<br />

to be hydrogenated with all the ligands but exceptionally<br />

ligand 7 (Table 4, entry 8). The alkyl enol carbamate 14<br />

is best hydrogenated by ligands 4a and 7, again leading<br />

to an inversion of enantioselectivity with ligand 7. In general<br />

the behaviour of substrates 12 and 14 is comparable<br />

to that of the previous enol carbamates.<br />

Table 4. Variations of the substituent in the olefin a<br />

3. Conclusions<br />

Monodentate phosphine ligands based on a 4,5-dihydro-<br />

3H-dinaphtho[2,1-c;1 0 ,2 0 -e]phosphepines structure 4 and 7<br />

have been success<strong>full</strong>y exploited in the rhodium-catalyzed<br />

asymmetric hydrogenation of various enol carbamates.<br />

The influences of different reaction parameters have been<br />

investigated. For the first time in this reaction a high<br />

enantioselectivity (up to 96% ee) was obtained in the presence<br />

of monodentate phosphines.<br />

4.1. General<br />

N<br />

O O<br />

12<br />

O O<br />

4. Experimental<br />

1 H, 13 C and 31 P NMR spectra were recorded on Bruker<br />

Spectrometer AVANCE 500, 400 and 300 ( 1 H: 500.13<br />

MHz, 400.13 MHz and 300.13 MHz; 13 C: 125.8 MHz,<br />

100.6 MHz and 75.5 MHz; 31 P: 162.0 MHz). The calibration<br />

N<br />

13<br />

N<br />

O O<br />

Entry Lig. 12a 13a 14a<br />

Conv.<br />

(%)<br />

ee b<br />

(%)<br />

Conv.<br />

(%)<br />

ee<br />

(%)<br />

14<br />

Conv.<br />

(%)<br />

ee<br />

(%)<br />

1 4a 94 50 (S) 99 48<br />

2 4b 10 12 (R)


of 1 H and 13 C spectra was carried out on solvent signals<br />

(d(CDCl3) = 7.25 and 77.0). The 31 P chemical shifts are<br />

referenced to 85% H3PO4. Mass spectra were recorded on<br />

an AMD 402 spectrometer. Optical rotations were<br />

measured on a Gyromat-HP polarimeter. IR spectra were<br />

recorded as KBr pellets or Nujol mulls on a Nicolet Magna<br />

550. All manipulations with air sensitive compounds were<br />

performed under argon atmosphere using standard<br />

Schlenk techniques. Toluene was distilled from sodium<br />

benzophenone ketyl under argon. Methanol was distilled<br />

from Mg under argon. Ethanol and 2-propanol were<br />

distilled from Na under argon. Methylene chloride was distilled<br />

from CaH2 under argon. Dimethyl sulfoxide (on<br />

molecular sieves) and [Rh(cod)2]BF4 were purchased from<br />

Fluka and used without further purifications. Ligands 4<br />

were synthesized according to our previously published<br />

protocols. 11<br />

4.2. Synthesis of ligands<br />

4.2.1. (S)-4-Phenyl-4,5-dihydro-3H-dinaphtho[2,1-c:10 ,20-e] phosphepine oxide 5a. Ligand 4a (1.5 g, 3.86 mmol) were<br />

dissolved in acetone (10 ml) and 7.7 ml of 10% sol. H2O2 (23 mmol) were added care<strong>full</strong>y, cooling with ice-bath.<br />

After few minutes stirring at room temperature, the solution<br />

was refluxed at 100 °C for 1 h. The solvent was evaporated<br />

and the resulting yellow oil taken up with CH2Cl2<br />

and dried over MgSO4. Purification by column chromatography<br />

(eluent: ethyl acetate/petroleum ether 4:1) yielded<br />

1.17 g (75%) of white solid. Mp: 167–168 °C. 1 HNMR<br />

(500 MHz, CDCl3) d = 8.02 (d, 1H, 3 J3,4 = 8.5 Hz, H-4);<br />

7.96, 7.95 (2d, 2H, 3 J5,6 = 8.2 Hz, 3J 0 0<br />

5 ;6 ¼ 8:2 Hz, H-5,<br />

H-50 ); 7.90 (d, 1H, 3J 0 0<br />

3 ;4 ¼ 8:5 Hz, H-40 ); 7.70 (d, 1H,<br />

3<br />

J3,4 = 8.5 Hz, H-3); 7.53–7.35 (m, 7H, H-6, H-60 , Ph);<br />

7.29–7.21 (m, 3H), 7.16 (d, 1H, 3 J = 8.5 Hz, H-7, H-70 ,<br />

H-8, H-80 ); 7.21 (d, 1H, 3J 0 0<br />

3 ;4 ¼ 8:5 Hz, H-30 ); 3.36 (dd,<br />

1H, 2 JH,H = 14.2 Hz, 2 JP,H = 8.2 Hz, H-11a); 3.32 (dd,<br />

1H, 2 JP,H = 22.5 Hz, 2 JH,H = 14.2 Hz, H-110a); 3.22 (dd,<br />

1H, 2 JH,H = 14.2 Hz, 2 JP,H = 13.0 Hz, H-11b); 3.21 (dd,<br />

1H, 2 JP,H = 16.0 Hz, 2 JH,H = 14.2 Hz, H-110b). 13 CNMR<br />

(125.8 MHz, CDCl3): d = 134.0 (d, J = 4.0 Hz), 133.4 (d,<br />

J = 4.0 Hz), 133.0 (d, J = 1.8 Hz), 132.9 (d, J = 1.8 Hz),<br />

132.4 (d, J = 2.5 Hz), 132.2 (d, J = 1.8 Hz), 130.4 (d,<br />

J = 8.2 Hz), 129.6 (d, J = 10.0 Hz, C-1, C-10 , C-2, C-20 ,<br />

C-9, C-90 , C-10, C-100 ); 132.1 (d, J = 2.5 Hz, p-Ph); 132.0<br />

(d, J = 88.0 Hz, i-Ph); 130.9 (d, J = 8.5 Hz, o-Ph); 129.4<br />

(d, J = 1.8 Hz, C-4); 128.6 (d, J = 1.8 Hz, C-40 ); 128.5 (d,<br />

J = 11.0 Hz, m-Ph); 128.5 (d, J = 3.5 Hz, C-3); 128.4,<br />

128.2 (C-5, C-50 ); 128.3 (d, J = 4.5 Hz, C-30 ); 127.1,<br />

126.7, 126.5, 126.3 (C-7, C-70 , C-8, C-80 ); 125.9 (C-6);<br />

125.6 (C-60 ); 37.4 (d, J = 65.0 Hz, C-110 ); 36.0 (d,<br />

J = 64.0 Hz, C-11).<br />

31 P NMR (121.5 MHz, CDCl3)<br />

d = 54.0. IR (KBr, cm 1 ): 3052 m; 3008 w; 2951 w; 1618<br />

w; 1592 w; 1508 m; 1435 m; 1405 m; 1359 w; 1328 w;<br />

1251 m; 1221 s; 1199 m; 1159 m; 1114 m; 1102 m; 1061<br />

w; 1026 w; 998 w; 973 w; 960 w; 931 w; 884 w; 866 w;<br />

836 s; 833 m; 820 s; 802 w; 770 m; 752m; 742 s; 726 w;<br />

713 w; 703 m; 694 m; 672 w; 661 m; 622 m; 586 w; 568<br />

w; 544 m; 523 m; 496 w; 470 m; 436 w; 424 w. MS (ESI):<br />

m/z (%) = 404 ([M + ], 100); 365 (6); 266 (34); 202 (4); 139<br />

(4); 73 (7). HRMS calcd for C28H21O1P: 404.13245, found:<br />

404.132628. ½aŠ 22<br />

D ¼þ79:0 (c 0.46, CHCl3). Retention time:<br />

S. Enthaler et al. / Tetrahedron: Asymmetry 18 (2007) 1288–1298 1293<br />

64.6 min (30 m HP Agilent Technologies 50-8-260/5-8-280/<br />

5-8-300/5).<br />

4.2.2. (S)-4-(4-Methoxy)phenyl-4,5-dihydro-3H-dinaphtho-<br />

[2,1-c:10 ,20-e]phosphepine oxide 5b. (S)-4-(4-Methoxy)phenyl-4,5-dihydro-3H-dinaphtho[2,1-c:10<br />

,20-e]phosphepine 4d (4.2 mmol) was dissolved in acetone (15 ml) and a mixture<br />

of water (1.1 ml) and 35% sol. H2O2 (5.0 mmol) was<br />

added care<strong>full</strong>y. After stirring for 4 h at room temperature,<br />

the solvent was evaporated and the residue dissolved in<br />

dichloromethane. The organic phase was washed with<br />

water, brine and dried over MgSO4. The solvent was<br />

removed to obtain an off-white foam. Yield: 99%. Mp:<br />

112–115 °C. 1 H NMR (300 MHz, CDCl3) d = 8.03–7.89<br />

(m, 4H); 7.69 (dd, 1H, J = 8.5 Hz, J = 1.2 Hz); 7.51–7.41<br />

(m, 2H); 7.40–7.31 (m, 2H, C6H4); 7.29–7.15 (m, 5H);<br />

6.89 (m, 2H, C6H4); 3.83 (s, 3H, OMe); 3.37–3.12<br />

(m, 4H, CH2). 13 C NMR (75.5 MHz, CDCl3) d = 162.5<br />

(d, J = 3.0 Hz, C6H4); 133.9 (d, J = 3.8 Hz); 133.5 (d,<br />

J = 3.8 Hz); 132.9 (d, J = 2.0 Hz); 132.7 (d, J = 2.0 Hz);<br />

132.7 (d, J = 10.2 Hz, C6H4); 132.4 (d, J = 2.5 Hz); 132.1<br />

(d, J = 2.0 Hz); 130.6 (d, J = 8.2 Hz); 129.8 (d,<br />

J = 9.5 Hz); 129.3 (d, J = 1.8 Hz); 128.6 (d, J = 1.8 Hz);<br />

128.4 (d, J = 3.8 Hz); 128.4; 128.3 (d, J = 4.5 Hz); 128.2;<br />

127.1; 126.7; 126.5; 126.3; 125.8; 125.6; 123.0 (d,<br />

J = 93.5 Hz, C6H4); 114.0 (d, J = 12.0 Hz, C6H4); 55.3<br />

(OMe); 37.7 (d, J = 66.0 Hz, CH2); 36.3 (d, J = 64.2 Hz,<br />

CH2). 31 P NMR (121.5 MHz, CDCl3) d = 54.1. IR (KBr,<br />

cm 1 ): 3053 m; 2957 w; 2836 w; 1596 s; 1569 w; 1503 s;<br />

1460 m; 1441 m; 1407 m; 1294 m; 1255 s; 1217 m; 1178 s;<br />

1117 s; 1027 m; 931 w; 871 w; 816 s; 744 m; 684 w; 660<br />

w; 623 w; 567 w; 544 m; 527 m; 496 w; 457 w; 419 w.<br />

MS (EI): m/z (%) = 434 ([M + ], 100); 282 (20); 279 (37);<br />

266 (32); 121 (27). HRMS calcd for C29H23O2P:<br />

434.14302, found: 434.142805. ½aŠ 22<br />

D ¼ 293 (c 0.25,<br />

CHCl3).<br />

4.2.3. (S,S,Sa)-3,5-Dimethyl-4-phenyl-4,5-dihydro-3H-dinaphtho[2,1-c:1<br />

0 ,2 0 -e]phosphepine oxide 6. Phosphepine<br />

oxide 5 (1.37 g, 3.4 mmol) is dissolved in 14 ml of THF,<br />

followed by the addition of 0.42 ml (6.8 mmol) of methyl<br />

iodide. To the well stirred solution, 22 ml of LDA<br />

(11 mmol, 0.5 M) are added slowly at room temperature.<br />

The dark red solution was stirred for 1 h and afterwards<br />

quenched with few drops of water. The solvents were evaporated<br />

under reduced pressure. The product was taken up<br />

with 50 ml of CH2Cl2, washed with 50 ml of H2O and the<br />

water phase extracted with CH2Cl2 (2 · 50 ml). The organic<br />

phases are dried over MgSO 4 and the solvents evaporated,<br />

yielding 1.45 g of light yellow foam. Purification by<br />

column chromatography on SiO2, eluting with acetone/<br />

petroleum ether 1:1 yielded 0.56 g (38%) of a white solid.<br />

Mp: 292–305 °C. 1 H NMR (500 MHz, CDCl3) d = 8.02<br />

(d, 1H, 3 J 3 0 ;4 0 ¼ 8:5 Hz, H-4 0 ); 7.97 (d, 1H, 3 J3,4 = 8.5 Hz,<br />

H-4); 7.95 (d, 1H, 3 J5,6 = 8.2 Hz, H-5); 7.92 (d, 1H,<br />

3 J 5 0 ;6 0 ¼ 8:2 Hz, H-5 0 ); 7.72–7.68 (m, 2H, o-Ph); 7.65 (d,<br />

1H, 3 J3,4 = 8.5 Hz, H-3); 7.55 (d, 1H, 3 J 3 0 ;4 0 ¼ 8:5 Hz, H-<br />

3 0 ); 7.50–7.39 (m, 5H, H-6, H-6 0 , m-, p-Ph); 7.24–7.18 (m,<br />

3H), 7.05 (d, 1H, 3 J = 8.5 Hz, H-7, H-7 0 , H-8, H-8 0 ); 3.62<br />

(m, 1H, 2 JP,H = 15.0 Hz, 3 J 11 0 ;12 0 ¼ 7:7 Hz, H-11 0 ); 3.43<br />

(dq, 1H, 2 JP,H = 22.0 Hz, 3 J11,12 = 7.7 Hz, H-11); 0.95<br />

(dd, 1H, 3 J P,H = 14.0 Hz, 3 J 11 0 ;12 0 ¼ 7:7 Hz, H-12 0 ); 0.61


1294 S. Enthaler et al. / Tetrahedron: Asymmetry 18 (2007) 1288–1298<br />

3 JP,H = 16.1 Hz,<br />

3 J11,12 = 7.7 Hz, H-12).<br />

13 C<br />

(dd, 1H,<br />

NMR (125.8 MHz, CDCl3): d = 135.2 (d, J = 3.2 Hz, C-<br />

10 ); 134.5 (d, J = 5.0 Hz, C-20 ); 134.0 (d, J = 7.0 Hz, C-2);<br />

133.9 (d, J = 3.0 Hz), 133.8 (d, J = 1.8 Hz), 133.4 (d,<br />

J = 1.8 Hz), 133.0 (d, J = 1.5 Hz), 132.8 (d, J = 1.5 Hz,<br />

C-1, C-9, C-90 , C-10, C-100 ); 133.2 (d, J = 83.0 Hz, i-Ph);<br />

131.7 (d, J = 2.5 Hz, p-Ph); 131.1 (d, J = 8.2 Hz, o-Ph);<br />

130.1 (d, J = 5.0 Hz, C-3); 129.3 (d, J = 6.8 Hz, C-30 );<br />

129.2 (C-4); 129.1 (C-40 ); 128.2, 128.0 (C-5, C-50 ); 128.2<br />

(d, J = 10.0 Hz, m-Ph); 127.0, 126.6, 126.5, 126.3 (C-7, C-<br />

70 , C-8, C-80 ); 126.1 (C-6); 125.7 (C-60 ); 46.4 (d,<br />

J = 59.5 Hz, C-11); 43.2 (d, J = 63.0 Hz, C-110 ); 17.0 (C-<br />

12); 14.6 (d, J = 4.5 Hz, C-120 ). 31 P NMR (121.5 MHz,<br />

CDCl3) d = 51.3. IR (KBr, cm 1 ): 3039 m; 2990 m; 2931<br />

m; 2874 w; 1618 w; 1594 m; 1568 w; 1504 m; 1455 w;<br />

1436 m; 1375 w; 1341 w; 1325 w; 1292 w; 1253 m; 1225<br />

w; 1184 m; 1166 s; 1112 m; 1092 m; 1057 m; 1028 w; 999<br />

w; 957 w; 915 w; 896 m; 871 w; 838 m; 820 s; 796 w; 768<br />

w; 756 m; 739 s; 711 m; 704 m; 689 m; 667 s; 632 w; 583<br />

w; 564 s; 534 m; 520 w; 509 w; 482 m; 459 m; 418 w; 403<br />

w. MS (ESI): m/z (%) = 432 ([M + ], 44); 417 (2); 388 (2);<br />

360 (1); 328 (2); 293 (15); 278 (100); 265 (15); 231 (4); 208<br />

(12); 179 (2); 145 (16); 132 (33); 77 (7); 47 (4). HRMS calcd<br />

for C30H25OP: 432.16375, found: 432.163562. ½aŠ 22<br />

D ¼þ93:5<br />

(c 0.25, CHCl3). Retention time: 58.6 min (30 m HP Agilent<br />

Technologies 50-8-260/5-8-280/5-8-300/5).<br />

4.2.4. (S,S,Sa)-3,5-Dimethyl-4-phenyl-4,5-dihydro-3H-dinaphtho[2,1-c:10<br />

,20-e]phosphepine 7. Phosphepine oxide 6<br />

(0.08 g, 0.2 mmol) was dissolved in dry toluene (6 ml)<br />

and triethylamine (0.2 ml, 1.4 mmol) was added to the<br />

solution. The mixture was cooled to 0 °C with an ice-bath<br />

and HSiCl3 (0.1 ml, 1 mmol) was added. After the addition<br />

was complete, the solution was refluxed for 4 h. The reaction<br />

was quenched by 5 ml basic aqueous solution, and<br />

the aqueous phase washed with toluene (2 · 5 ml). The<br />

organic phase was dried over Na2SO4 and the solvent<br />

evaporated under reduced pressure. Yield 70 mg (91%).<br />

Mp: 212 °C. 1 H NMR (300 MHz, CDCl3) d = 8.04–7.90<br />

(m, 4H); 7.73–7.63 (m, 3H); 7.55 (d, 1H, J = 8.5 Hz);<br />

7.51–7.38 (m, 5H); 7.26–7.16 (m, 3H); 7.06 (d, 1H, J =<br />

8.5 Hz); 3.62 (m, 1H, 2 JP,H = 15.5 Hz, 3 JH,Me = 7.5 Hz,<br />

CHMe); 3.43 (dq, 1H, 2 JP,H = 22.0 Hz, 3 JH,Me = 7.5 Hz,<br />

CHMe); 0.95 (dd, 3H, 3 JP,H = 13.9 Hz, 3 JH,Me = 7.5 Hz,<br />

Me); 0.61 (dd, 3H, 3 JP,H = 16.0 Hz, 3 JH,Me = .5 Hz, Me).<br />

13<br />

C NMR (75.5 MHz, CDCl3) d = 141.2 (d, J = 2.5 Hz);<br />

138.5 (d, J = 1.0 Hz); 138.5 (d, J = 25.8 Hz); 134.7 (d,<br />

J = 5.8 Hz); 134.3; 133.9; 133.86 (d, J = 2.8 Hz); 132.8 (d,<br />

J = 19.5 Hz); 132.7; 132.2; 128.9 (d, J = 3.0 Hz); 128.8;<br />

128.6 (d, J = 1.0 Hz); 128.4; 128.3 (d, J = 3.2 Hz); 128.2;<br />

128.1; 127.9; 126.6; 126.5; 126.0; 126.0; 125.2; 125.1; 39.3<br />

(d, J = 20.0 Hz, CHMe); 36.6 (d, J = 20.0 Hz, CHMe);<br />

22.3 (d, J = 35.5 Hz, Me); 14.0 (d, J = 3.5 Hz, Me). 31 P<br />

NMR d = 31.46 (s). MS (ESI): m/z (%): 416 ([M + ], 100),<br />

360 (10), 265 (38), 149 (9). HRMS: calcd for C30H25P<br />

416.16884, found: 416.167881. ½aŠ 25<br />

D ¼þ87:5 (c 0.2,<br />

CHCl3).<br />

4.3. General synthesis of enol carbamate derivates 8–14 14,5m<br />

A solution of sodium hydride (0.11 mol) in dry dimethyl<br />

sulfoxide (200 ml) was stirred for 2 h at 50 °C under an<br />

atmosphere of argon. The grey suspension was added dropwise<br />

to the corresponding ketone (0.1 mol) in 25 ml of dimethyl<br />

sulfoxide at room temperature. The orange<br />

solution was stirred for 15 min at room temperature and<br />

then cooled to 10 °C. The addition of the corresponding<br />

carbamoyl chloride (0.11 mol) in dimethyl sulfoxide<br />

(0.11 mol) was carried out in 30 min while maintaining<br />

the temperature at 10 °C. After stirring overnight at room<br />

temperature, the solution was care<strong>full</strong>y quenched with<br />

water (250 ml), extracted with n-hexane or heptane<br />

(3 · 250 ml), washed with brine (250 ml) and dried over<br />

MgSO4 or Na2SO4. The solvents were removed in vacuo<br />

and the obtained yellow oil purified by column chromatography<br />

or crystallization.<br />

4.3.1. 1-Phenylvinyl N,N-diethylcarbamate 8. Purification<br />

by column chromatography (eluent: ethylacetate/n-hexane<br />

1<br />

1:10) yielded a yellow oil. Yield: 18%. H NMR<br />

(300 MHz, CDCl3) d = 7.44–7.39 (m, 2H, Ph); 7.30–7.19<br />

(m, 3H, Ph); 5.35 (d, 1H, J = 2.0 Hz, @CH2); 4.96 (d,<br />

1H, J = 2.0 Hz, @CH2); 3.35 (q, 2H, J = 7.0 Hz, CH2); 3.25 (q, 2H, J = 7.0 Hz, CH2); 1.20 (t, 3H, J = 7.0 Hz,<br />

13<br />

CH3); 1.10 (t, 3H, J = 7.0 Hz, CH3). C NMR<br />

(75.5 MHz, CDCl3) d = 153.8; 153.4; 135.2; 128.6; 128.4;<br />

124.8; 101.3; 42.0; 41.7; 14.3; 13.3. IR (KBr, cm 1 ):<br />

3059 w; 3027 w; 2975 m; 2934 m; 2876 w; 1719 s; 1641<br />

m; 1600 w; 1577 w; 1495 m; 1474 m; 1457 m; 1420 s;<br />

1380 m; 1351 w; 1314 w; 1255 s; 1225 m; 1158 s; 1097 m;<br />

1076 m; 1050 m; 1027 w; 981 w; 902 w; 869 m; 787 m;<br />

771 m; 759 m; 703 m; w; 638 w; 578 w. MS (ESI): m/z<br />

(%) = 219 ([M + ], 9); 103 (12); 100 (100); 77 (16); 72 (72);<br />

44 (16). HRMS calcd for C13H17O2N: 219.1254, found:<br />

219.1253. Retention time: 17.8 min (30 m HP Agilent<br />

Technologies 50-8-260/5-8-280/5-8-300/5).<br />

4.3.2. 1-Phenylvinyl N,N-dimethylcarbamate 9. Purification<br />

by column chromatography on silica gel eluting with<br />

hexane/ethyl acetate 4:1 yielded 6.4 g of an orange oil.<br />

Yield: 34%. 1 H NMR (300 MHz, CDCl3) d = 7.48–7.28<br />

(m, 5H, Ph); 5.41 (1H, d, J = 2.0 Hz, @CH 2); 5.02 (d,<br />

1H, J = 2.0 Hz, @CH2); 3.10 (s, 3H, CH3); 2.96 (s, 3H,<br />

CH3). 13 C NMR (75.5 MHz, CDCl3) d = 154.5; 153.4;<br />

135.1; 128.7; 128.5; 124.9; 101.6; 36.7; 36.4. IR (KBr,<br />

cm 1 ): 3084 w, 3058 w, 3023 w, 2934 m, 1958 w, 1724 s,<br />

1643 m, 1601 w, 1577 m, 1494 s, 1445 s, 1391 s, 1313 m,<br />

1295 m, 1262 s, 1168 s, 1098 m, 1076 m, 1030 m, 928 m,<br />

875 m, 858 m, 772 s, 759 m, 708 m, 692 m, 641 w, 581 w.<br />

MS (ESI): m/z (%) = 191 ([M + ], 12); 103 (4); 91 (4); 77<br />

(7); 72 (100); 51 (5). HRMS calcd for C11H13O2N:<br />

191.09408, found: 191.09370. Retention time: 16.8 min<br />

(30 m HP Agilent Technologies 50-8-260/5-8-280/5-8-300/<br />

5).<br />

4.3.3. 1-Phenylvinyl N-methyl N-phenylcarbamate 10.<br />

Purification by column chromatography on silica gel eluting<br />

with n-hexane/ethyl acetate 1:1 followed by crystallization<br />

with acetone/diethylether/n-hexane 1:2:4. Yield: 25%.<br />

Mp: 56–59 °C. 1 H NMR (300 MHz, CDCl3) d = 7.42–<br />

7.26 (m, 10H, Ph); 5.39 (1H, d, J = 1.7 Hz, @CH 2); 5.06<br />

(d, 1H, J = 1.7 Hz, @CH2); 3.38 (br s, 3H, CH3). 13 C<br />

NMR (75.5 MHz, CDCl3) d = 153.6; 153.4; 143.0; 134.9;<br />

129.2; 128.8; 128.4; 126.7 (br); 126.0 (br); 125.0; 101.6;


38.3. IR (KBr, cm 1 ): 3762 w; 3407 w; 3119 w; 3085w; 3067<br />

w; 3037 w; 2975 w; 2937 w; 2557 w; 1966w; 1891 w; 1813 w;<br />

1717 s; 1643 m; 1592 m; 1578 w; 1493 s; 1446 m; 1420 m;<br />

1370 s; 1294 m; 1257 s; 1180 w; 1146 s; 1088 m; 1072 m;<br />

1050 w; 1025 m; 1004 w; 981 m; 916 w; 875 m; 840 w;<br />

827 w; 771 m; 754 m; 705 s; 687 m; 602 w; 588 m; 563 w;<br />

529 w; 429 w; 408 m. MS (ESI): m/z (%) = 253 ([M + ], 4);<br />

134 (100); 119 (4); 106 (37); 91 (6); 77 (38); 65 (4); 51<br />

(10); 39 (3). HRMS calcd for C16H15O2N: 253.10973,<br />

found: 253.10977. Retention time: 23.1 min (30 m HP Agilent<br />

Technologies 50-8-260/5-8-280/5-8-300/5).<br />

4.3.4. 1-Phenylvinyl N,N-diphenylcarbamate 11. Crystallization<br />

from acetone/diethylether/n-hexane 1:2:5 yielded<br />

white-greenish crystals. Yield: 38%. Mp: 72–75 °C. 1 H<br />

NMR (300 MHz, CDCl3) d = 7.26–7.21 (m, 15H, Ph);<br />

5.40 (1H, d, J = 2.2 Hz, @CH 2); 5.14 (1H, d, J = 2.2 Hz,<br />

@CH2). 13 C NMR (75.5 MHz, CDCl3) d = 153.2; 152.5;<br />

142.2; 134.7; 129.1; 128.8; 128.4; 126.8; 126.5; 125.0;<br />

101.6. IR (KBr, cm 1 ): 3088 w; 3057 m; 1952 w; 1879 w;<br />

1801 w; 1721 s; 1646 m; 1590 m; 1492 s; 1451 m; 1384 w;<br />

1347 s; 1328 m; 1307 m; 1293 m; 1256 s; 1201 s; 1184 m;<br />

1172 m; 1097 m; 1078 m; 1044 m; 1032 m; 1024 m; 913<br />

w; 870 m; 836 w; 770 s; 761 s; 694 s; 638 m; 620 w; 597<br />

m; 566 w; 530 m; 512 m; 443 w; 412 w. MS (ESI): m/z<br />

(%) = 315 ([M + ], 13); 196 (100); 168 (46); 103 (17); 91 (4);<br />

77 (22); 51 (7). HRMS calcd for C 21H 17O 2N: 315.12538,<br />

found: 315.12538. Retention time: 28.2 min (30 m HP Agilent<br />

Technologies 50-8-260/5-8-280/5-8-300/5).<br />

4.3.5. (Z)-1-Phenylprop-1-enyl N,N-diethylcarbamate 12. 18<br />

Colourless oil was obtained after column chromatography<br />

(eluent: ethyl acetate/n-hexane 1:10). Yield: 38%. 1 H NMR<br />

(500 MHz, CDCl3) d = 7.40 (m, 2H, o-Ph); 7.30 (m, 2H, m-<br />

Ph); 7.23 (m, 2H, p-Ph); 5.85 (q, 1H, 3 J1,2 = 7.0 Hz, H-2);<br />

3.49, 3.36 (2q, 4H, H-5, H-5 0 ); 1.74 (d, 3H, 3 J1,2=7.0 Hz,<br />

H-1); 1.29, 1.17 (2t, 6H, 3 J 5,6 = 7.0 Hz, 3 J 5 0 ;6 0 ¼ 7:0 Hz,<br />

H-6, H-6 0 ). 13 C NMR (125.8 MHz, CDCl3) d = 153.4<br />

(C@O); 147.3 (C–O); 136.0 (i-Ph); 128.3 (m-Ph); 127.6 (p-<br />

Ph); 124.3 (o-Ph); 112.4 (C-2); 42.1, 41.7 (C-5, C-5 0 );<br />

14.4, 13.4 (C-6, C-6 0 ); 11.4 (C-1). IR (KBr, cm 1 ): 3058<br />

w; 2975 m; 2934 m; 2875 w; 1717 s; 1673 m; 1600 w;<br />

1577 w; 1495 m; 1474 m; 1458 m; 1420 s; 1380 m; 1351<br />

w; 1313 m; 1258 s; 1225 m; 1158 s; 1116 m; 1097 m; 1061<br />

m; 1033 w; 1006 m; 972 w; 951 w; 926 w; 851 w; 793 w;<br />

753 s; 692 m; 652 w; 637 w; 574 w. MS (ESI): m/z<br />

(%) = 233 ([M + ], 6); 115 (9); 105 (8); 100 (100); 77 (12);<br />

72 (52); 44 (12). HRMS calcd for C14H19O2N: 233.14103,<br />

found: 233.14101. Retention time: 19.3 min (30 m HP Agilent<br />

Technologies 50-8-260/5-8-280/5-8-300/5). Retention<br />

time (HPLC): 8.72 min (eluent: n-hexane/ethanol 98:2;<br />

flow: 1.0 ml/min).<br />

4.3.6. 1H-Inden-3-yl N,N-diethyl carbamate 13. An<br />

orange oil was obtained after column chromatography<br />

(eluent: ethyl acetate/n-hexane 1:10). Yield: 58%. 1 H<br />

NMR (300 MHz, CDCl3) d = 7.34–7.10 (m, 4H); 6.19 (t,<br />

1H, J = 2.3 Hz, CH); 3.41–3.27 (m, 4H, CH2CH3); 3.29<br />

(d, 2H, J = 2.3 Hz, CH 2CH); 1.22–1.08 (m, 6H, CH 3).<br />

13 C NMR (75.5 MHz, CDCl3) d = 153.0; 149.4; 141.9;<br />

139.7; 126.1; 125.4; 124.0; 117.8; 113.6; 42.3; 42.1; 34.8;<br />

14.3; 13.4. IR (KBr, cm 1 ): 3074 w; 3024 w; 2975 m;<br />

S. Enthaler et al. / Tetrahedron: Asymmetry 18 (2007) 1288–1298 1295<br />

2934 m; 2890 m; 1726 s; 1653 w; 1616 m; 1605 m; 1577<br />

m; 1474 m; 1459 m; 1420 s; 1380 m; 1361 m; 1316 m;<br />

1267 s; 1235 m; 1223 m; 1204 m; 1172 s; 1156 s; 1118 m;<br />

1098 m; 1076 m; 1017 w; 973 m; 953 m; 932 m; 915 w;<br />

844 w; 761 s; 717 m; 635 w; 593 w; 553 w; 511 w; 467 w;<br />

412 w. MS (ESI): m/z (%) = 231 ([M + ], 6); 131 (10); 115<br />

(6); 103 (11); 100 (100); 77 (14); 72 (62); 44 (17). HRMS<br />

calcd for C 14H 17O 2N: 231.12538; Found: 231.124840.<br />

Retention time: 21.3 min (30 m HP Agilent Technologies<br />

50-8-260/5-8-280/5-8-300/5).<br />

4.3.7. 3,3-Dimethylbut-1-en-2-yl N,N-diethyl carbamate 14.<br />

A colourless oil was obtained after column chromatography<br />

(eluent: ethyl acetate/n-hexane 1:3). Yield: 27%. 1 H<br />

NMR (400 MHz, CDCl3) d = 4.76 (d, 1H, J = 1.9 Hz,<br />

@CH2); 4.64 (d, 1H, J = 1.9 Hz, @CH2); 3.31 (q, 4H,<br />

J = 7.0 Hz, CH2CH3); 1.16–1.12 (br, 6H, CH2CH3); 1.10 (s, 9H, CH3).<br />

13 C NMR (100.6 MHz, CDCl3)<br />

d = 163.0; 154.1; 96.2; 42.0; 41.6; 36.2; 27.9; 14.2; 13.4.<br />

IR (KBr, cm 1 ): 3125 w; 2973 s; 2936 m; 2876 m; 1720 s;<br />

1653 m; 1555 w; 1506 w; 1474 m; 1461 m; 1419 s; 1380<br />

m; 1362 m; 1317 m; 1264 s; 1225 m; 1148 s; 1097 m; 1054<br />

m; 979 m; 938 w; 859 m; 786 m; 756 m; 703 w; 655 w;<br />

593 w; 488 w. MS (ESI): m/z (%) = 199 ([M + ], 1); 118<br />

(1); 100 (100); 85 (2); 72 (50); 67 (3); 55 (5); 44 (15). HRMS<br />

calcd for C11H21O2N: 199.15668, found: 199.155983.<br />

Retention time: 11.5 min (30 m HP Agilent Technologies<br />

50-8-260/5-8-280/5-8-300/5).<br />

4.4. General procedure for the catalytic hydrogenation of<br />

enol carbamates<br />

The catalyst was generated in situ by stirring [Rh(cod) 2]-<br />

BF4 (0.0024 mmol) and the corresponding 4,5-dihydro-<br />

3H-dinaphthophosphepines ligand (0.005 mmol) in 1.0 ml<br />

of solvent for a period of 10 min and afterwards transferring<br />

via syringe into the autoclave. A solution of the enol<br />

carbamate (0.24 mmol) and 1.0 ml solvent was transferred<br />

via syringe into the autoclave. Then, the autoclave was<br />

charged with hydrogen and stirred at the required temperature.<br />

After a predetermined time the hydrogen was released<br />

and the reaction mixture passed through a short<br />

plug of silica gel. The conversion and enantioselectivity<br />

were measured by GC and HPLC without further<br />

modifications.<br />

4.4.1. 1-Phenylethyl N,N-diethylcarbamate 8a. Colourless<br />

oil. 1 H NMR (300 MHz, CDCl3) d = 7.31–7.11 (m, 5H,<br />

Ph); 5.73 (d, 1H, J = 6.5 Hz, CH); 3.20 (q, 4H,<br />

J = 7.0 Hz, CH2); 1.44 (d, 3H, J = 6.5 Hz, CH3); 1.02 (t,<br />

6H, J = 7.0 Hz, CH3). 13 C NMR (75.5 MHz, CDCl3)<br />

d = 155.1; 142.6; 128.2; 127.3; 125.7; 72.6; 41.5; 41.1;<br />

22.7; 14.0; 13.4. IR (KBr, cm 1 ): 3087 w; 3064 w; 3033<br />

w; 2977 s; 2933 m; 2875 w; 1700 s; 1630 m; 1586 w; 1539<br />

w; 1495 m; 1476 s; 1457 s; 1425 s; 1379 m; 1316 m; 2174<br />

s; 1227 m; 1210 m; 1172 s; 1069 s; 1030 m; 1010 m; 998<br />

m; 973 m; 940 w; 911 w; 877 w; 787 m; 766 m; 700 s; 630<br />

w; 576 w; 535 m. MS (ESI): m/z (%) = 221 ([M + ], 2); 105<br />

(100); 100 (5); 77 (16); 72 (5); 58 (7); 51 (6); 44 (8). HRMS<br />

calcd for C13H19O2N: 221.14103, found: 221.140593.<br />

½aŠ 23<br />

D ¼ 146:1 (c 0.5, CH2Cl2/MeOH, 96% ee (S)). Conversions<br />

were determined by GC (Agilent Technologies, 30 m,


1296 S. Enthaler et al. / Tetrahedron: Asymmetry 18 (2007) 1288–1298<br />

50–300 °C, (50-8-260/5-8-280/5-8-300/5)) and ees by HPLC<br />

(Whelk (R,R), (S)-8a 8.29 min and (R)-8a 28.7 min (eluent:<br />

n-hexane/ethanol 99:1; flow: 1.0 ml/min)).<br />

4.4.2. (S)-1-Phenylethyl N,N-diethylcarbamate 8a. To a<br />

solution of (S)-1-phenylethanol (3.3 mmol) in THF<br />

(10 ml) was added NaH (3.4 mmol). The solution was stirred<br />

for 30 min at room temperature under an argon atmosphere.<br />

The solution was cooled to 0 °C and N,N-diethyl<br />

carbamoyl chloride (3.4 mmol) in THF (5 ml) was added.<br />

After stirring for 2 h at room temperature the reaction<br />

was quenched with water. The mixture was extracted with<br />

ethylacetate/n-hexane (1:1), washed with brine and dried<br />

over Na 2SO 4. The solvents were removed in vacuum and<br />

the crude oil was purified by flash column chromatography<br />

(ethylacetate/n-hexane 1:1) to yield a colourless oil (yield:<br />

56%). The analytical data are in agreement with 8a<br />

obtained by the hydrogenation protocol.<br />

4.4.3. 1-Phenylethyl N,N-dimethylcarbamate 9a. Oil. 1 H<br />

NMR (400 MHz, CDCl3) d = 7.34–7.22 (m, 5H, Ph); 5.78<br />

(q, 1H, J = 6.6 Hz, CH); 2.93 (br s, 3H, CH3); 2.88 (br s,<br />

3H, CH3); 1.51 (d, 3H, J = 6.6 Hz, CH3). 13 C NMR<br />

(100.6 MHz, CDCl 3) d = 156.0; 142.7; 128.5; 127.6; 125.9;<br />

73.1; 36.4; 35.9; 22.9. IR (KBr, cm 1 ): 3033 w; 2980 w;<br />

2932 w; 1704 s; 1495 m; 1450 m; 1393 m; 1273 w; 1191 s;<br />

1069 m; 1030 w; 1007 w; 995 w; 894 w; 844 w; 766 m;<br />

700 m; 638 w; 569 w; 532 w. MS (ESI): m/z (%) = 193<br />

([M + ], 6); 134 (13); 121 (1); 105 (100); 90 (1); 77 (13); 63<br />

(1); 51 (4). HRMS calcd for C 11H 15O 2N: 193.10973, found:<br />

193.109345. ½aŠ 22<br />

D ¼ 2:9 (c 0.3, CHCl3, 75% ee (S)). Conversions<br />

were determined by NMR and ees by HPLC<br />

(Chiralpak AD-H, (R)-9a 23.6 min and (S)-9a 31.3 min,<br />

eluent: n-hexane/ethanol 99:1; flow: 1.0 ml/min). The absolute<br />

configuration was assigned by analogy.<br />

4.4.4. 1-Phenylethyl N-methyl N-phenylcarbamate 10a.<br />

Oil. 1 H NMR (300 MHz, CDCl3) d = 7.35–7.16 (m, 10H,<br />

Ph); 5.83 (q, 1H, J = 6.5 Hz, CH); 3.28 (s, 3H, CH3);<br />

1.47 (d, 3H, J = 6.5 Hz, CHCH 3). 13 C NMR (75.5 MHz,<br />

CDCl3) d = 155.0; 143.4; 142.3; 128.8; 128.4; 127.6; 125.9<br />

(br); 125.8; 125.7 (br); 73.8; 37.7; 23.0. IR (KBr, cm 1 ):<br />

3064 w; 3033 w; 3980 m; 2932 w; 1706 s; 1598 m; 1496 s;<br />

1453 m; 1437 m; 1422 m; 1374 s; 1327 m; 1299 s; 1277 s;<br />

1210 m; 1159 s; 1113 m; 1063 s; 1029 m; 1010 m; 997 m;<br />

974 m; 912 w; 885 w; 819 w; 762 s; 698 s; 665 w; 625 w;<br />

599 w; 544 m. MS (ESI): m/z (%) = 255 ([M + ], 1); 211<br />

(2); 196 (7); 151 (1); 134 (2); 105 (100); 91 (1); 77 (20); 65<br />

(2); 51 (7). HRMS calcd for C16H17NO2: 255.12538, found:<br />

255.125229. ½aŠ 22<br />

D ¼þ19:9 (c 0.26, CHCl 3, 65% ee (S)).<br />

Conversions were determined by NMR and ees by HPLC<br />

(Chiralcel OD-H, (R)-10a 25.4 min and (S)-10a 47.2 min,<br />

eluent: n-hexane/ethanol 99:1; flow: 1.0 ml/min). The absolute<br />

configuration was assigned by analogy.<br />

4.4.5. 1-Phenylethyl N,N-diphenylcarbamate 11a. Crys-<br />

1<br />

tals. Mp: 76–78 °C. H NMR (300 MHz, CDCl3)<br />

d = 7.32–7.14 (m, 15H, Ph), 5.88 (q, 1H, J = 6.5 Hz,<br />

13<br />

CH), 1.47 (d, 3H, J = 6.5 Hz, CH3). C NMR<br />

(75.5 MHz, CDCl3) d = 154.1; 142.6; 141.9; 128.9; 128.4;<br />

127.7; 127.0; 126.1; 125.9; 74.3; 22.9. IR (KBr, cm 1 ):<br />

3064 w; 3033 m; 2984 m; 2934 w; 1702 s; 1591 s; 1491 s;<br />

1450 m; 1369 s; 1343 s; 1325 s; 1299 s; 1281 s; 1220 s;<br />

1208 s; 1178 m; 1159 w; 1128 w; 1058 s; 1048 s; 1023 s;<br />

1008 m; 993 m; 951 w; 913 w; 901 w; 860 m; 837 w; 814<br />

w; 781 m; 764 s; 758 s; 693 s; 672 m; 642 w; 622 w; 603<br />

w; 548 s; 516 m; 492 w. MS (ESI): m/z (%) = 317 ([M + ],<br />

2); 273 (2); 258 (2); 196 (1); 169 (41); 139 (1); 105 (100);<br />

77 (12); 51 (5). HRMS calcd for C21H19O2N: 317.14103,<br />

found: 317.141098. ½aŠ 21<br />

D ¼ 16:8 (c 0.33, CHCl 3,76%ee<br />

(R)). Conversions were determined by NMR and ees by<br />

HPLC (Chiralcel OD-H, (R)-11a 6.8 min and (S)-11a<br />

9.6 min, eluent: n-hexane/ethanol 99.5:0.5; flow: 1.0 ml/<br />

min). The absolute configuration was assigned by analogy.<br />

4.4.6. 1-Phenylpropyl N,N-diethylcarbamate 12a. Colour-<br />

1<br />

less oil. Yield: 98%. H NMR (400 MHz, CDCl3)<br />

d = 7.28–7.16 (m, 5H, Ph); 5.54 (1H, dd, J = 6.4 Hz,<br />

J = 7.2 Hz, CHCH2); 3.23 (br, 4H, NCH2); 1.83 (m, 2H,<br />

CHCH2CH3); 1.06 (br, 6H, NCH2CH3); 0.83 (t, 3H,<br />

J = 7.4 Hz, CHCH2CH3). 13 C NMR (100.6 MHz, CDCl3)<br />

d = 155.5; 141.6; 128.3; 127.5; 126.4; 77.8; 41.8; 41.3;<br />

29.9; 14.3; 13.6; 9.9. IR (KBr, cm 1 ): 2972 m; 2934 m;<br />

2877 w; 1700 s; 1475 m; 1457 m; 1424 m; 1380 w; 1273 s;<br />

1226 w; 1171 s; 1063 m; 987 m; 903 w; 763 w; 700 m; 528<br />

w. MS (ESI): m/z (%) = 235 ([M + ], 9); 162 (15); 119 (54);<br />

100 (12); 91 (100); 77 (7). HRMS calcd for C14H19O2N:<br />

235.15668, found: 235.157242. ½aŠ 23<br />

D ¼ 143:7 (c0.5, CH2Cl2/MeOH, 50% ee). Conversions and ees were determined<br />

by GC (Retention time: 17.9 min, 30 m HP Agilent<br />

Technologies 50–300 °C: 50-8-260/5-8-280/5-8-300/5) and<br />

HPLC (Whelk (R,R), n-hexane/ethanol 98:2, flow 1.0 ml/<br />

min, (S)-12a 4.84 min and (R)-12a 10.10 min).<br />

4.4.7. 3,3-Dimethylbutan-2-yl N,N-diethylcarbamate 14a.<br />

1<br />

Oil. H NMR (300 MHz, CDCl3) d = 4.56 (q, 1H,<br />

J = 6.4 Hz, CH); 3.25 (br d, 4H, CH2); 1.12 (d, 3H,<br />

J = 6.4 Hz, CH3); 1.10 (t, 6H, CH2CH3); 0.90 (s, 9H,<br />

CH3). 13 C NMR (75.5 MHz, CDCl3) d = 155.9; 78.1;<br />

41.6; 34.4; 25.9; 15.3; 14.2. IR (KBr, cm 1 ): 2971 s; 2935<br />

m; 2874 w; 1696 s; 1653 w; 1559 w; 1540 w; 1506 w; 1475<br />

m; 1457 m; 1423 s; 1395 w; 1378 m; 1365 w; 1316 w;<br />

1275 m; 1227 w; 1210 w; 1174 m; 1076 m; 1007 w; 972 w;<br />

893 w; 824 w; 783 w; 768 w; 744 w; 697 w; 617 w; 537 w.<br />

MS (ESI): m/z (%) = 201 ([M + ], 7); 186 (3); 144 (2); 116<br />

(16); 100 (100); 85 (44); 72 (25); 58 (21); 43 (46). HRMS<br />

calcd for C11H23O2N: 201.17233, found: 201.172292.<br />

½aŠ 22<br />

D ¼þ135 (c 0.04, CHCl3/MeOH, 67% ee). Conversions<br />

and ees were determined by GC (50m Chiraldex b-PH, 50.0<br />

m · 250 lm · 0.25 lm, 100/25-4-180/5, ( )-14a 25.6 min<br />

and (+)-14a 26.0 min).<br />

4.5. X-ray crystallographic studies of 5a, 5b and 6<br />

Data were collected with a STOE-IPDS diffractometer<br />

using graphite-monochromated Mo Ka radiation. The<br />

structures were solved by direct methods 19 and refined by<br />

<strong>full</strong>-matrix least-squares techniques against F 2 . 20 XP<br />

(BRUKER AXS) was used for graphical representations.<br />

The crystallographic data have been deposited with the<br />

Cambridge Crystallographic Data Centre as supplementary<br />

publication no. CCDC 641493–641495. Copies of the<br />

data can be obtained free of charge on application to<br />

http://www.ccdc.cam.ac.uk/data_request/cif.


4.5.1. Compound 5a. Space group P3 121, trigonal, a =<br />

11.174(2), c = 32.520(7) A ˚ , V = 3516(1) A ˚ 3 , Z =6,<br />

qcalcd = 1.230 g cm 3 , 18,941 reflections measured, 3621<br />

were independent of symmetry, of which 2550 were observed<br />

(I >2r(I)), R1 = 0.055, wR 2 (all data) = 0.145, 281<br />

parameters, Flack parameter x = 0.05(18).<br />

4.5.2. Compound 5b. Space group P2 1, monoclinic, a =<br />

7.985(2), b = 11.407(2), c = 12.625(3) A ˚ , b = 98.27(3)°,<br />

V = 1138.0(4) A ˚ 3 , Z =2, qcalcd = 1.262 g cm 3 , 15,617<br />

reflections measured, 4452 were independent of symmetry,<br />

of which 2645 were observed (I >2r(I)), R1 = 0.034, wR 2<br />

(all data) = 0.059, 289 parameters, Flack parameter<br />

x = 0.02(8).<br />

4.5.3. Compound 6. Space group P21, monoclinic, a =<br />

9.445(2), b = 16.801(3), c = 14.081(3) A ˚ , b = 91.75(3)°,<br />

V = 2233.4(8) A ˚ 3 , Z =4, qcalcd = 1.292 g cm 3 , 13,676<br />

reflections measured, 7244 were independent of symmetry,<br />

of which 5881 were observed (I >2r(I)), R1 = 0.047, wR 2<br />

(all data) = 0.118, 577 parameters, Flack parameter x =<br />

0.07(10).<br />

Acknowledgements<br />

We thank Mrs. M. Heyken, Mrs. S. Buchholz and Dr. C.<br />

Fischer (all Leibniz-Institut für Katalyse e.V. an der <strong>Universität</strong><br />

<strong>Rostock</strong>) for excellent technical and analytical<br />

assistance. Dr. B. Hagemann is grate<strong>full</strong>y thanked for the<br />

synthesis of ligand 4e. Generous financial support from<br />

the state of Mecklenburg-Western Pomerania and the<br />

BMBF as well as the Deutsche Forschungsgemeinschaft<br />

(Leibniz-price) are grate<strong>full</strong>y acknowledged.<br />

References<br />

1. (a) Lennon, I. C.; Ramsden, J. A. Org. Process Res. Dev.<br />

2005, 9, 110–112; (b) Hawkins, J. M.; Watson, T. J. N.<br />

Angew. Chem. 2004, 116, 3286–3290; (c) Noyori, R.;<br />

Ohkuma, T. Angew. Chem. 2001, 113, 40–75; Miyagi, M.;<br />

Takehara, J.; Collet, S.; Okano, K. Org. Process Res. Dev.<br />

2000, 4, 346–348; (d) Comprehensive Asymmetric Catalysis;<br />

Jacobsen, E. N., Pfaltz, A., Yamamoto, H., Eds.; Springer:<br />

Berlin, 1999; (e) Noyori, R. Asymmetric Catalysis in Organic<br />

Synthesis; Wiley: New York, 1994.<br />

2. (a) Anastas, P. T.; Kirchhoff, M. M. Acc. Chem. Res. 2002,<br />

35, 686–694; (b) Anastas, P. T.; Kirchhoff, M. M.; Williamson,<br />

T. C. Appl. Catal. A: Gen. 2001, 221, 3–13.<br />

3. Lagasse, F.; Kagan, H. B. Chem. Pharm. Bull. 2000, 48, 315–<br />

324.<br />

4. Komarov, I. V.; Börner, A. Angew. Chem. 2001, 113, 1237–<br />

1240.<br />

5. (a) Arnold, L. A.; Imobos, R.; Manoli, A.; de Vries, A. H.<br />

M.; Naasz, R.; Feringa, B. Tetrahedron 2000, 56, 2865–2878;<br />

(b) van den Berg, M.; Minnaard, A. J.; Schudde, E. P.; van<br />

Esch, J.; de Vries, A. H. M.; de Vries, J. G.; Feringa, B. L. J.<br />

Am. Chem. Soc. 2000, 122, 11539–11540; (c) Minnaard, A. J.;<br />

van den Berg, M.; Schudde, E. P.; van Esch, J.; de Vries, A.<br />

H. M.; de Vries, J. G.; Feringa, B. L. Chim. Oggi 2001, 19,<br />

12–13; (d) Peña, D.; Minnaard, A. J.; de Vries, J. G.; Feringa,<br />

B. L. J. Am. Chem. Soc. 2002, 124, 14552–14553; (e) van den<br />

S. Enthaler et al. / Tetrahedron: Asymmetry 18 (2007) 1288–1298 1297<br />

Berg, M.; Haak, R. M.; Minnaard, A. J.; de Vries, A. H. M.;<br />

de Vries, J. G.; Feringa, B. L. Adv. Synth. Catal. 2002, 344,<br />

1003–1007; (f) van den Berg, M.; Minnaard, A. J.; Haak, R.<br />

M.; Leeman, M.; Schudde, E. P.; Meetsma, A.; Feringa, B.<br />

L.; de Vries, A. H. M.; Maljaars, C. E. P.; Willans, C. E.;<br />

Hyett, D.; Boogers, J. A. F.; Henderickx, H. J. W.; de Vries,<br />

J. G. Adv. Synth. Catal. 2003, 345, 308–323; (g) Peña, D.;<br />

Minnaard, A. J.; de Vries, A. H. M.; de Vries, J. G.; Feringa,<br />

B. L. Org. Lett. 2003, 5, 475–478; (h) Jiang, X.; van den Berg,<br />

M.; Minnaard, A. J.; Feringa, B. L.; de Vries, J. G.<br />

Tetrahedron: Asymmetry 2004, 15, 2223–2229; (i) Eelkema,<br />

R.; van Delden, R. A.; Feringa, B. L. Angew. Chem., Int. Ed.<br />

2004, 43, 5013–5016; (j) Duursma, A.; Boiteau, J.-G.; Lefort,<br />

L.; Boogers, J. A. F.; de Vries, A. H. M.; de Vries, J. G.;<br />

Minnaard, A. J.; Feringa, B. L. J. Org. Chem. 2004, 69,<br />

8045–8052; (k) Lefort, L.; Boogers, J. A. F.; de Vries, A. H.<br />

M.; de Vries, J. G. Org. Lett. 2004, 6, 1733–1735; (l)<br />

Bernsmann, H.; van den Berg, M.; Hoen, R.; Minnaard, A.<br />

J.; Mehler, G.; Reetz, M. T.; de Vries, J. G.; Feringa, B. L. J.<br />

Org. Chem. 2005, 70, 943–951; (m) Panella, L.; Feringa, B.<br />

L.; de Vries, J. G.; Minnaard, A. J. Org. Lett. 2005, 7, 4177–<br />

4180; (n) Hoen, R.; Boogers, J. A. F.; Bernsmann, H.;<br />

Minnaard, A. J.; Meetsma, A.; Tiemersma-Wegman, T. D.;<br />

de Vries, A. H. M.; de Vries, J. G.; Feringa, B. L. Angew.<br />

Chem., Int. Ed. 2005, 44, 4209–4212; (o) Panella, L.;<br />

Aleixandre, A. M.; Kruidhof, G. J.; Robertus, J.; Feringa,<br />

B. L.; de Vries, J. G.; Minnaard, A. J. J. Org. Chem. 2006,<br />

71, 2026–2036.<br />

6. (a) Reetz, M. T.; Sell, T. Tetrahedron Lett. 2000, 41, 6333–<br />

6336; (b) Reetz, M. T.; Mehler, G. Angew. Chem., Int. Ed.<br />

2000, 39, 3889–3891; (c) Reetz, M. T.; Mehler, G.; Meiswinkel,<br />

G.; Sell, T. Tetrahedron Lett. 2002, 43, 7941–7943; (d)<br />

Reetz, M. T.; Mehler, G.; Meiswinkel, G.; Sell, T. Angew.<br />

Chem., Int. Ed. 2003, 42, 790–793; (e) Reetz, M. T.; Mehler,<br />

G. Tetrahedron Lett. 2003, 44, 4593–4596; (f) Reetz, M. T.;<br />

Mehler, G.; Meiswinkel, A. Tetrahedron: Asymmetry 2004,<br />

15, 2165–2167; (g) Reetz, M. T.; Li, X. G. Tetrahedron 2004,<br />

60, 9709–9714; (h) Reetz, M. T.; Ma, J.-A.; Goddard, R.<br />

Angew. Chem., Int. Ed. 2005, 44, 412–414; (i) Reetz, M. T.;<br />

Fu, Y.; Meiswinkel, A. Angew. Chem. 2006, 118, 1440–1443;<br />

(j) Reetz, M. T.; Li, X. Synthesis 2005, 3183–3185; (k) Reetz,<br />

M. T.; Bondarev, O. G.; Gais, H.-J.; Bolm, C. Tetrahedron<br />

Lett. 2005, 46, 5643–5646; (l) Reetz, M. T.; Meiswinkel, A.;<br />

Mehler, G.; Angermund, K.; Graf, M.; Thiel, W.; Mynott,<br />

R.; Blackmond, D. G. J. Am. Chem. Soc. 2005, 127, 10305–<br />

10313; (m) Reetz, M. T.; Li, X. Angew Chem., Int. Ed. 2005,<br />

44, 2959–2962; (n) Reetz, M. T.; Surowiec, M. Heterocycles<br />

2006, 67, 567–574.<br />

7. (a) Claver, C.; Fernandez, E.; Gillon, A.; Heslop, K.; Hyett,<br />

D. J.; Martorell, A.; Orpen, A. G.; Pringle, P. G. Chem.<br />

Commun. 2000, 961–962; (b) Martorell, A.; Naasz, R.;<br />

Feringa, B. L.; Pringle, P. G. Tetrahedron: Asymmetry<br />

2001, 12, 2497–2499.<br />

8. (a) Hu, A.-G.; Fu, Y.; Xie, J.-H.; Zhou, H.; Wang, L.-X.;<br />

Zhou, Q.-L. Angew. Chem. 2002, 114, 2454–2456; (b) Fu, Y.;<br />

Guo, X.-X.; Zhu, S.-F.; Hu, A.-G.; Xie, J.-H.; Zhou, Q.-L. J.<br />

Org. Chem. 2004, 69, 4648–4655; (c) Fu, Y.; Hou, G.-H.; Xie,<br />

J.-H.; Xing, L.; Wang, L.-X.; Zhou, Q.-L. J. Org. Chem.<br />

2004, 69, 8157–8160.<br />

9. (a) Gladiali, S.; Dore, A.; Fabbri, D.; De Lucchi, O.;<br />

Manassero, M. Tetrahedron: Asymmetry 1994, 5, 511–514;<br />

(b) Gladiali, S.; Dore, A.; Fabbri, D.; De Lucchi, O.; Valle,<br />

G. J. Org. Chem. 1994, 59, 6363–6371; (c) Gladiali, S.;<br />

Fabbri, D. Chem. Ber./Recl. Trav. Chim. Pays-Bas 1997, 130,<br />

543–554; (d) Charruault, L.; Michelet, V.; Taras, R.; Gladiali,<br />

S.; Genet, J.-P. Chem. Commun. 2004, 850–851; (e) Zanoni,<br />

G.; Gladiali, S.; Marchetti, A.; Piccinini, P.; Tredici, I.;<br />

Vidari, G. Angew. Chem., Int. Ed. 2004, 43, 846–849; (f)


1298 S. Enthaler et al. / Tetrahedron: Asymmetry 18 (2007) 1288–1298<br />

Alberico, E.; Nieddu, I.; Taras, R.; Gladiali, S.. Helv. Chim.<br />

Acta 2006, 89, 1716–1729.<br />

10. (a) Chi, Y.; Zhang, X. Tetrahedron Lett. 2002, 43, 4849–4852;<br />

(b) Tang, W.; Wang, W.; Chi, Y.; Zhang, X. Angew. Chem.,<br />

Int. Ed. 2003, 42, 3506–3509; (c) Chi, Y.; Zhou, Y.-G.; Zhang,<br />

X. J. Org. Chem. 2003, 68, 4120–4122; (d) Xiao, D.; Zhang,<br />

X. Angew. Chem., Int. Ed. 2001, 40, 3425–3428; (e) Xiao, D.;<br />

Zhang, Z.; Zhang, X. Org. Lett. 1999, 1, 1679–1681.<br />

11. (a) Junge, K.; Oehme, G.; Monsees, A.; Riermeier, T.;<br />

Dingerdissen, U.; Beller, M. Tetrahedron Lett. 2002, 43,<br />

4977–4980; (b) Junge, K.; Oehme, G.; Monsees, A.; Riermeier,<br />

T.; Dingerdissen, U.; Beller, M. J. Organomet. Chem.<br />

2003, 675, 91–96; (c) Junge, K.; Hagemann, B.; Enthaler, S.;<br />

Spannenberg, A.; Michalik, M.; Oehme, G.; Monsees, A.;<br />

Riermeier, T.; Beller, M. Tetrahedron: Asymmetry 2004, 15,<br />

2621–2631; (d) Junge, K.; Hagemann, B.; Enthaler, S.;<br />

Oehme, G.; Michalik, M.; Monsees, A.; Riermeier, T.;<br />

Dingerdissen, U.; Beller, M. Angew. Chem. 2004, 116,<br />

5176–5179; Angew. Chem., Int. Ed. 2004, 43, 5066–5069; (e)<br />

Hagemann, B.; Junge, K.; Enthaler, S.; Michalik, M.;<br />

Riermeier, T.; Monsees, A.; Beller, M. Adv. Synth. Catal.<br />

2005, 347, 1978–1986; (f) Enthaler, S.; Hagemann, B.; Junge,<br />

K.; Erre, G.; Beller, M. Eur. J. Org. Chem. 2006, 2912–<br />

2917.<br />

12. (a) Wilson, J. E.; Fu, G. C. Angew. Chem., Int. Ed. 2006, 45,<br />

1426–1429; (b) Wurz, R. P.; Fu, G. C. J. Am. Chem. Soc.<br />

2005, 127, 12234–12235; (c) Vedejs, E.; Daugulis, O.; Diver,<br />

S. T. J. Org. Chem. 1996, 61, 430–431.<br />

13. Kasák, P.; Mereiter, K.; Widhalm, M. Tetrahedron: Asymmetry<br />

2005, 16, 3416–3426.<br />

14. (a) Olofson, R. A.; Cuomo, J.; Bauman, B. A. J. Org. Chem.<br />

1978, 43, 2073–2075; (b) Jiang, X.-b.; van den Berg, M.;<br />

Minnaard, A. J.; Feringa, B. L.; de Vries, J. G. Tetrahedron:<br />

Asymmetry 2004, 15, 2223–2229.<br />

15. Institute of Automation (IAT), Richard Wagner Str. 31/H. 8,<br />

18119 <strong>Rostock</strong>-Warnemünde, Germany.<br />

16. Tsukamoto, M.; Yoshimura, M.; Tsuda, K.; Kitamura, M.<br />

Tetrahedron 2006, 62, 5448–5453.<br />

17. The reaction is stopped after 1 h and a black precipitate<br />

(rhodium-black) is formed.<br />

18. During the synthesis the Z-isomer is formed exclusively,<br />

which was characterized by NMR. In a NOESY spectrum a<br />

correlation was observed between the ortho-hydrogen of the<br />

phenyl group and the olefinic hydrogen.<br />

19. Sheldrick, G. M. SHELXS-97, SHELXS Program of Crystal Structure<br />

Solution; University of Göttingen: Germany, 1997.<br />

20. Sheldrick, G. M. SHELXL-97, SHELXL Program of Crystal Structure<br />

Refinement; University of Göttingen: Germany, 1997.


6. Transfer hydrogenation<br />

6.1. Homogeneous iron catalyst in reduction chemistry<br />

6.1.1. An environmentally benign process for the hydrogenation of ketones<br />

with homogeneous iron catalysts<br />

Stephan Enthaler, Bernhard Hagemann, Giulia Erre, Kathrin Junge, and Matthias Beller*,<br />

Chem. Asian. J. 2006, 1, 598–604.<br />

Contributions<br />

The BuP(Ad)2 ligand was synthesized by Dr. R. Jackstell.<br />

100


FULL PAPERS<br />

DOI: 10.1002/asia.200600105<br />

An Environmentally Benign Process for the Hydrogenation of Ketones with<br />

Homogeneous Iron Catalysts<br />

Stephan Enthaler, Bernhard Hagemann, Giulia Erre, Kathrin Junge, and<br />

Matthias Beller* [a]<br />

Abstract: Iron complexes generated<br />

in situ catalyze homogeneously the<br />

transfer hydrogenation of aliphatic and<br />

aromatic ketones by utilizing 2-propanol<br />

as a hydrogen donor in the presence<br />

of base. The influence of different<br />

reaction parameters on the catalytic activity<br />

is investigated in detail by apply-<br />

Introduction<br />

Catalysis is a key technology for the advancement of green<br />

chemistry, specifically for waste prevention, decreasing<br />

energy consumption, achieving high atom efficiency and<br />

generating advantageous economics. [1] There is an increasing<br />

interest in substituting toxic and expensive late transition<br />

metals by readily available and less-toxic metals. In this<br />

regard, the use of iron catalysts is especially desirable. [2] So<br />

far, homogeneous iron catalysts have been success<strong>full</strong>y applied<br />

for various C C coupling reactions such as Friedel–<br />

Crafts-type reactions, olefin polymerizations, cross-couplings,<br />

cycloadditions, and substitution reactions. [3] However,<br />

much less is known in the area of industrially important catalytic<br />

reductions. Here, comparatively few Fe-catalyzed hydrogenations<br />

have been established, mainly for the reduction<br />

of olefins [4] and nitro compounds. [5] Clearly, the quest<br />

for practical hydrogenation catalysts based on Fe complexes<br />

constitutes a major challenge for the development of more<br />

sustainable reductions.<br />

Recently, we became interested in applying homogeneous<br />

Fe catalysts with respect to C H functionalization reactions<br />

[a] S. Enthaler, B. Hagemann, G. Erre, Dr. K. Junge, Prof. Dr. M. Beller<br />

Leibniz-Institut für Katalyse e.V. an der <strong>Universität</strong> <strong>Rostock</strong><br />

Albert-Einstein-Str. 29a<br />

18059 <strong>Rostock</strong> (Germany)<br />

Fax: (+ 49) 381-1281-5000<br />

E-mail: matthias.beller@ifok-rostock.de<br />

ing a three-component catalyst system<br />

composed of an iron salt, 2,2’:6’,2’’-terpyridine,<br />

and PPh 3. The scope and limi-<br />

Keywords: homogeneous catalysis ·<br />

iron · ketones · ligand effects ·<br />

transfer hydrogenation<br />

tations of the described catalyst is<br />

shown in the reduction of 11 different<br />

ketones. In most cases, high conversion<br />

and excellent chemoselectivity are obtained.<br />

Mechanistic studies indicate a<br />

monohydride reaction pathway for the<br />

homogeneous iron catalyst.<br />

of arenes. [6] Based on that work and our ongoing research in<br />

hydrogenation chemistry, [7] we started to explore Fe catalysts<br />

for transfer hydrogenations [8] of carbonyl compounds.<br />

To the best of our knowledge, only iron carbonyls [9,10] or<br />

complexes that contain tetradentate aminophosphines [11] or<br />

phosphines [12] have been described for the transfer hydrogenations<br />

of ketones and a,b-unsaturated carbonyl compounds.<br />

In the latter case, mainly hydrogenation of the<br />

olefin occurred. [9,12]<br />

Our goal is to develop a practical Fe hydrogenation catalyst<br />

system, which should be easy to prepare and tunable.<br />

Thus, the use of commercially available Fe complexes in<br />

combination with two different ligands (phosphines and<br />

amines) instead of tetradentate ligands seems to be a useful<br />

approach. Based on this idea, we report herein the application<br />

of new three-component iron catalysts prepared in situ<br />

based on iron salts, 2,2’:6’,2’’-terpyridine (terpy), and PPh 3.<br />

These catalysts give excellent yield and selectivity in the reduction<br />

of aromatic and aliphatic ketones to alcohols.<br />

Results and Discussion<br />

As a starting point, 2-propanol-based transfer hydrogenation<br />

of acetophenone (1) was examined with Fe catalysts in the<br />

presence of combinations of nitrogen and phosphorus ligands.<br />

In exploratory experiments, terpy and PPh 3 were<br />

used as ligands. Typically, the precatalyst was prepared<br />

in situ by stirring a solution of [Fe 3(CO) 12] (0.3 mol %),<br />

terpy (1 mol %), and PPh 3 (1 mol %) in 2-propanol (1.0 mL)<br />

598 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Asian J. 2006, 1, 598 – 604


for 16 h at 65 8C. Initially, we investigated the influence of<br />

terpy and PPh 3 in the presence of [Fe 3(CO) 12], which can be<br />

easily handled without special precautions (Table 1). To our<br />

Table 1. Transfer hydrogenation of acetophenone (1) with<br />

ACHTUNGTRENUNG[Fe 3(CO) 12]/terpy/PPh 3 catalyst. [a]<br />

Entry Terpy [b]<br />

PPh 3 [b]<br />

Yield [%] [c]<br />

Selectivity [%] [d]<br />

1 – – 23 > 99<br />

2 1 1 78 > 99<br />

3 1 – 18 > 99<br />

4 – 1 6 > 99<br />

5 – 2 21 > 99<br />

6 – 10 29 > 99<br />

7 5 1 27 > 99<br />

8 1 5 24 > 99<br />

[a] Reaction conditions: in situ catalyst (0.0038 mmol) ([Fe3(CO) 12]<br />

(0.0013 mmol), terpy (0.0038 mmol), PPh3 (0.0038 mmol), 2-propanol<br />

(2.0 mL) for 16 h at 658C), iPrONa (0.019 mmol), 5 min at 100 8C, then<br />

addition of 1 (0.38 mmol), 7 h at 100 8C. [b] Ligand to Fe ratio. [c] Yield<br />

was determined by GC (50 m Lipodex E, 95–200 8C) with diglyme as internal<br />

standard (yield is equivalent to conversion). [d] Selectivity refers<br />

to chemoselectivity.<br />

delight, a 1:1 mixture of terpy and PPh 3 gave an active catalyst<br />

for the test reaction that was superior to all other combinations<br />

(Table 1, entry 2). Notably, low catalytic activity<br />

was observed without the use of any ligands (Table 1,<br />

entry 1). Increasing the ligand concentration resulted in a<br />

significant decrease in activity (Table 1, entries 7 and 8). Importantly,<br />

the catalyst systems differ mainly with respect to<br />

reactivity, as the chemoselectivity was excellent in all cases.<br />

Next, the influence of base and the iron precatalyst was<br />

investigated in more detail (Table 2). The best results were<br />

obtained with [Fe 2(CO) 9] and [Fe 3(CO) 12] in the presence of<br />

catalytic amounts of sodium 2-propylate or sodium tert-bu-<br />

ACHTUNGTRENUNGtylate (Table 2, entries 1, 6, 16). Surprisingly, the most commonly<br />

used bases for transfer hydrogenation, such as<br />

NaOH, KOH, and tBuOK, showed only low activity in this<br />

model reaction (Table 2, entries 3–5). Furthermore, different<br />

inorganic bases such as K 2CO 3, Cs 2CO 3, and K 3PO 4<br />

(Table 2, entries 7–9) as well as nitrogen-containing organic<br />

bases, for instance, pyridine, NEt 3, NACHTUNGTRENUNG(iPr) 2Et, DBU, and<br />

DABCO (Table 2, entries 10–14), did not prove to be effective.<br />

As expected, no transfer of hydrogen was observed in<br />

the absence of base (Table 2, entry 15).<br />

To improve the catalytic system, various iron sources with<br />

different oxidation states (0, +2, and +3) were tested. Besides<br />

[Fe 2(CO) 9] and [Fe 3(CO) 12], FeCl 2 (Table 2, entry 20)<br />

also produced reasonable conversion. To facilitate the formation<br />

of active iron hydride complexes, we tested [Et 3NH]-<br />

ACHTUNGTRENUNG[HFe(CO) 4] [13] as precatalyst, but only limited conversion<br />

was detected (Table 2, entry 19).<br />

Following these results, we focused on the nature of the ligands.<br />

The results in Table 3 indicate no improvement of<br />

Table 2. Influence of different bases and iron sources in the Fe-catalyzed<br />

transfer hydrogenation of 1. [a]<br />

Entry Iron source Base Yield [%] [b]<br />

1 ACHTUNGTRENUNG[Fe3(CO) 12] iPrONa 78<br />

2 ACHTUNGTRENUNG[Fe3(CO) 12] LiOH 2<br />

3 ACHTUNGTRENUNG[Fe3(CO) 12] NaOH < 1<br />

4 ACHTUNGTRENUNG[Fe3(CO) 12] KOH < 1<br />

5 ACHTUNGTRENUNG[Fe3(CO) 12] tBuOK 12<br />

6 ACHTUNGTRENUNG[Fe3(CO) 12] tBuONa 76<br />

7 ACHTUNGTRENUNG[Fe3(CO) 12] K2CO3 3<br />

8 ACHTUNGTRENUNG[Fe3(CO) 12] Cs2CO3 3<br />

9 ACHTUNGTRENUNG[Fe3(CO) 12] K3PO4 < 1<br />

10 ACHTUNGTRENUNG[Fe3(CO) 12] pyridine 1<br />

11 ACHTUNGTRENUNG[Fe3(CO) 12] NEt3 2<br />

12 ACHTUNGTRENUNG[Fe3(CO) 12] NACHTUNGTRENUNG(iPr) 2Et < 1<br />

13 ACHTUNGTRENUNG[Fe3(CO) 12] DBU < 1<br />

14 ACHTUNGTRENUNG[Fe3(CO) 12] DABCO < 1<br />

15 ACHTUNGTRENUNG[Fe3(CO) 12] – 0<br />

16 ACHTUNGTRENUNG[Fe2(CO) 9] iPrONa 84<br />

17 [Fe(CO) 5] iPrONa 2<br />

18 ACHTUNGTRENUNG[CpFe(CO) 2I] iPrONa 3<br />

19 ACHTUNGTRENUNG[Et3NH]ACHTUNGTRENUNG[HFe(CO) 4] iPrONa 11<br />

20 FeCl2 iPrONa 45<br />

21 FeCl3·xH2O iPrONa < 1<br />

22 FeBr2 iPrONa 9<br />

23 FeSO4·7H2O iPrONa 9<br />

24 FeACHTUNGTRENUNG(acac) 2 iPrONa 17<br />

25 FeACHTUNGTRENUNG(acac) 3 iPrONa 3<br />

[a] Reaction conditions: in situ catalyst (0.0038 mmol) ([Fe3(CO) 12]<br />

(0.0013 mmol), terpy (0.0038 mmol), PPh3 (0.0038 mmol), 2-propanol<br />

(2.0 mL) for 16 h at 65 8C), base (0.019 mmol), 5 min at 100 8C, then addition<br />

of 1 (0.38 mmol), 7 h at 100 8C. [b] Yield was determined by GC<br />

(50 m Lipodex E, 95–200 8C) with diglyme as internal standard (yield is<br />

equivalent to conversion). acac = Acetylacetonate, Cp = cyclopentadienyl,<br />

DABCO=1,4-diazabicycloACHTUNGTRENUNG[2.2.2]octane,<br />

undec-7-ene.<br />

DBU= 1,8-diazabicycloACHTUNGTRENUNG[5.4.0]-<br />

conversion when PPh 3 was substituted by other phosphorus<br />

ligands. Variation in the substitution pattern of PPh 3 with<br />

electron-donating (Table 3, entry 2) or electron-withdrawing<br />

groups (Table 3, entries 3–5) as well as more-basic and sterically<br />

hindered phosphines (Table 3, entries 6–9) decreased<br />

the yield of 1-phenylethanol (2). We also applied diphosphine<br />

ligands in the model reaction. Good activity was obtained<br />

with 1,1-bis-(diphenylphosphanyl)methane (dppm)<br />

and 1,2-bis(diphenylphosphanyl)ethane (dppe) (Table 3, entries<br />

12–13). Next, we explored the nature of the nitrogencontaining<br />

ligand. To our surprise, substituted terpyridines<br />

such as 4’-chloro-2,2’:6’,2’’-terpyridine (3), 6,6’-dibromo-<br />

2,2’:6,6’-terpyridine (4), and 4,4’,4’’-tri-tert-butyl-2,2’:6,2’’-terpyridine<br />

(5) led to a significant decrease in alcohol formation<br />

(Table 4, entries 1–4). Similarly, the application of structurally<br />

related N,N’,N’’-ligands 6 and 7 showed no pronounced<br />

activity (Table 4, entries 5 and 6). However, in the<br />

presence of pyridine (3 mol %) and triACHTUNGTRENUNGphenylphosphine, a<br />

reasonable yield of 2 (54%) was obtained. Clearly, Fe/pyridine/PPh<br />

3 represents one of the least demanding homogenous<br />

transfer-hydrogenation catalysts around. Furthermore,<br />

Chem. Asian J. 2006, 1, 598 – 604 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemasianj.org 599


FULL PAPERS<br />

Table 3. Influence of phosphorus ligands in the Fe-catalyzed transfer<br />

ACHTUNGTRENUNGhydrogenation of 1. [a]<br />

Entry Phosphine Conversion [%] [b]<br />

1 PPh 3 78<br />

2 P(p-MeO-C 6H 4) 3 13<br />

3 P(p-Me-C 6H 4) 3 22<br />

4 PACHTUNGTRENUNG(p-F-C 6H 4) 3 23<br />

5 P(3,4-CF 3-C 6H 3) 3 5<br />

6 PCy 3 11<br />

7 PACHTUNGTRENUNG(tBu) 3 32<br />

8 34<br />

9 31<br />

10 PACHTUNGTRENUNG(OPh) 3 8<br />

11 PACHTUNGTRENUNG(OiPr) 3 7<br />

12 Ph2PCH2PPh2 64<br />

13 Ph2PACHTUNGTRENUNG(CH2) 2PPh2 50<br />

14 Ph2PACHTUNGTRENUNG(CH2) 4PPh2 3<br />

15 Ph2PACHTUNGTRENUNG(CH2) 6PPh2 3<br />

[a] Reaction conditions: in situ catalyst (0.0038 mmol) ([Fe3(CO) 12]<br />

(0.0013 mmol), terpy (0.0038 mmol), phosphorus ligand (0.0038 mmol), 2propanol<br />

(2.0 mL) for 16 h at 658C), iPrONa (0.019 mmol), 5 min at<br />

100 8C, then addition of 1 (0.38 mmol), 7 h at 100 8C. [b] Yield was determined<br />

by GC (50 m Lipodex E, 95–200 8C) with diglyme as internal standard<br />

(yield is equivalent to conversion). Cy = cyclohexyl.<br />

the combination of ( )-sparteine and triphenylphosphine<br />

provided an active catalyst system (Table 4, entries 11–12).<br />

Next, the optimized general protocol for transfer hydrogenations<br />

was applied to aromatic and aliphatic ketones. Here,<br />

both [Fe 3(CO) 12]/terpy/PPh 3 and FeCl 2/terpy/PPh 3 were<br />

tested for the reduction of 10 different ketones (Table 5).<br />

Acetophenone, 4-chloroacetophenone, 2-methoxyaceto-<br />

ACHTUNGTRENUNGphenone, and propiophenone were hydrogenated in excellent<br />

yield and selectivity (92–99 %) (Table 5, entries 1, 2, 5,<br />

6). Acetophenones with electron-donating substituents in<br />

the para position gave good but somewhat lower yields (75–<br />

83 %). A chloro substituent in the a position to the carbonyl<br />

group proved to be problematic and deactivated both catalysts<br />

(Table 5, entry 7). Aliphatic ketones are more challenging<br />

substrates than aromatic ketones, but they also react in<br />

excellent yield (95–99 %).<br />

Notably, similar activities for the reduction of ketones<br />

with regard to our [Fe 3(CO) 12]-based system were reported<br />

when other transition-metal carbonyl complexes, for instance,<br />

[Ru 3(CO) 12]-based catalysts, were applied. [14] The<br />

[Fe 3(CO) 12]- and FeCl 2-based catalysts showed no significant<br />

difference in productivity. Hence, we assume the formation<br />

of a similar active species. Indeed, the two catalyst systems<br />

produced similar conversion curves (Figure 1). At the start<br />

of the reaction, we observe in both cases an induction<br />

M. Beller et al.<br />

Figure 1. Conversion–time behavior of the different precatalysts containing<br />

[Fe 3(CO) 12] and FeCl 2. Reaction conditions: in situ catalyst<br />

(0.0038 mmol) ([Fe 3(CO) 12] (0.0013 mmol) or FeCl 2 (0.0038 mmol), terpy<br />

(0.0038 mmol), and PPh 3 (0.0038 mmol) in 2-propanol (2.0 mL), 16 h at<br />

658C), iPrONa (0.38 mmol), 5 min at 100 8C, then addition of 1<br />

(0.76 mmol), reaction at 100 8C. Conversion was determined by GC (50 m<br />

Lipodex E, 95-200 8C) with diglyme as internal standard (conversion is<br />

equivalent to yield). ^=[Fe 3(CO) 12]/terpy/PPh 3, & = FeCl 2/terpy/PPh 3.<br />

period of nearly one hour, whereby the FeCl 2 system<br />

showed a slightly lower reaction rate, probably due to the<br />

slower elimination of chlorides and the change of oxidation<br />

state. The induction period can be shortened by increasing<br />

the catalyst preformation time. Thus, when the precatalyst<br />

([Fe 3(CO) 12]/terpy/PPh 3) was treated with a base for one<br />

hour instead of 5 min at the reaction temperature, an increase<br />

in conversion into 1-phenylACHTUNGTRENUNGethanol from 18 % to<br />

32 % in the first hour was recorded.<br />

Next, we focused our attention on the reaction mechanism.<br />

To exclude the formation of heterogeneous Fe catalysts,<br />

[15] a large excess of Hg(0) was added to the well-stirred<br />

reaction mixture after the reaction had proceeded for one<br />

hour, so that the “real” catalyst should be formed. [16, 17] No<br />

significant suppression of the reaction rate in the reduction<br />

of 1 was observed (73 % yield after 7 h), whereas a positive<br />

poisoning of the catalyst should have led to approximately<br />

18 % of 2 (see also Figure 1). In another experiment, the reaction<br />

was carried out under standard conditions and filtered<br />

through celite after one hour. [18] The filtrate was allowed<br />

to react for a further six hours. Thereafter, the applied<br />

celite was stirred with fresh 2-propanol, sodium 2-propylate,<br />

and 1 under reaction conditions for another six<br />

hours. The results obtained showed no suppression of reaction<br />

rate for the filtrate (92 %), whereas no conversion was<br />

detected (< 1 %) with the celite system. Consequently, both<br />

experiments indicate a definite homogeneous catalyst.<br />

Various methods ( 1 H, 31 P, and 13 C NMR and IR spectroscopy<br />

and MS) were used for the characterization of the<br />

active catalyst species. Unfortunately, the structural composition<br />

of the catalyst is still unclear. No evidence for an Fe<br />

H species was detected by 1 H NMR spectroscopy after reaction<br />

of the precatalyst with base (5 equiv). [19] The 31 P NMR<br />

spectrum of the precatalyst in [D 4]MeOH or [D 8]iPrOH<br />

showed three singlets with a ratio of 20:1:10 at 5.3 ppm<br />

600 www.chemasianj.org 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Asian J. 2006, 1, 598 – 604


Ketone Hydrogenation with Homogeneous Iron Catalysts<br />

Table 4. Variation of nitrogen ligands in the Fe-catalyzed transfer<br />

ACHTUNGTRENUNGhydrogenation of 1. [a]<br />

Entry Ligand Ligand/metal Conversion [%] [b]<br />

1 terpy 1 78<br />

2 3 1 23<br />

3 4 1 28<br />

4 5 1 14<br />

5 6 1 13<br />

6 7 1 7<br />

7 2,2’-bipyridine 1 8<br />

8 2,2’-bipyridine 2 6<br />

9 tmeda [c]<br />

1 18<br />

10 tmeda [c]<br />

2 13<br />

11 ( )-sparteine 1 43 [d]<br />

12 ( )-sparteine 2 38 [d]<br />

13 pyridine 3 54<br />

14 pyridine 6 5<br />

[a] Reaction conditions: in situ catalyst (0.0038 mmol) ([Fe3(CO) 12]<br />

(0.0013 mmol), nitrogen ligand (0.0038 mmol), PPh3 (0.0038 mmol), 2propanol<br />

(2.0 mL) for 16 h at 658C), iPrONa (0.019 mmol), 5 min at<br />

100 8C, then addition of 1 (0.38 mmol), 7 h at 100 8C. [b] Yield was determined<br />

by GC (50 m Lipodex E, 95–200 8C) with diglyme as internal standard<br />

(yield is equivalent to conversion). [c] tmeda = N,N,N’,N’-tetrameth-<br />

ACHTUNGTRENUNGylethylendiamine. [d] A racemic mixture of 2 was detected.<br />

(free PPh 3) and two unknown signals at 32.9 and<br />

71.3 ppm. [20] The addition of 5 equivalents of base (sodium<br />

2-propylate) with respect to iron and stirring of the mixture<br />

for 5 min at 100 8C did not affect the chemical shift and<br />

ratio of the observed 31 P NMR signals. [19] Mass spectrometric<br />

investigations also gave no clear information about the composition<br />

of the precatalyst, because indications for a number<br />

of possible candidates or fragments of labile complexes<br />

were detected, such as compounds containing terpy, PPh 3,<br />

CO, and Fe in a ratio of 1:1:1(2):1 or terpy, CO, and Fe in a<br />

ratio of 1:3:1. The utilization of [Fe 3(CO) 12] represents a potential<br />

option for IR spectroscopy. Hence, we recorded the<br />

IR spectra of our precatalyst in a solution of 2-propanol.<br />

The activation of the precatalyst by sodium 2-propylate<br />

(10 equiv) for 5 min at 100 8C resulted in the http://<br />

www.dict.cc/?s =disappearance of absorption signals at 1889<br />

and 1718 cm 1 . Addition of 1 to the activated precatalyst led<br />

to the http://www.dict.cc/?s =disappearance of the signal at<br />

1654 cm 1 and emergence of a signal at 1679 cm 1 . Interestingly,<br />

a similar behavior was described by Gao and co-workers<br />

when they followed the formation of the transfer-hydrogenation<br />

catalyst composed of [Fe 3(CO) 12] and tetradentate<br />

aminophosphines in 2-propanol in the presence of base. [11]<br />

Although the nature of the active Fe H species remains<br />

unclear, we turned our attention to the mechanism of the<br />

hydride transfer. To exclude a radical-type reduction, the reaction<br />

of cyclopropyl phenyl ketone was examined in more<br />

detail (“radical clock” substrate) (Table 5, entry 8). In the<br />

presence of [Fe 3(CO) 12]/terpy/PPh 3 catalyst, the corresponding<br />

cyclopropyl phenyl alcohol 14 was detected by 1 H NMR<br />

spectroscopy in >99 % purity. There was apparently no radical-induced<br />

reduction, because no opening of the cyclopro-<br />

ACHTUNGTRENUNGpyl ring occurred. [21] Consequently, a radical-reduction<br />

mechanism promoted by sodium alkoxides, whereby the<br />

transition metal plays a marginal role, can also be excluded.<br />

[22]<br />

In general, for transition-metal-catalyzed transfer hydrogenation,<br />

two mechanisms are accepted: direct hydrogen<br />

transfer via formation of a six-membered cyclic transition<br />

state composed of metal and hydrogen donor and acceptor,<br />

and the hydridic route, which is subdivided into two pathways,<br />

the monohydride and the dihydride mechanism<br />

(Scheme 1). More specifically, the formation of monohydride<br />

metal complexes promote an exclusive hydride transfer<br />

from carbon (donor) to carbonyl carbon (acceptor)<br />

(Scheme 1, pathway A), whereas a hydride transfer from<br />

carbon (donor) to carbonyl carbon (acceptor) as well as carbonyl<br />

oxygen (acceptor) was proposed for the formation of<br />

dihydride metal complexes (Scheme 1, pathway B). [8a,c,e] Evidence<br />

for both pathways (hydridic route) were determined<br />

by various researchers when investigating the hydride transfer<br />

catalyzed by metal complexes of, for example, Ru, Rh,<br />

or Ir. [23] So far nothing is known with respect to iron catalysts<br />

in transfer hydrogenations.<br />

To rule out an exchange of hydrogen atoms, for example,<br />

by C H activation, we investigated the transfer hydrogenation<br />

with a completely deuterated donor molecule. [24] The<br />

[Fe 3(CO) 12]/terpy/PPh 3 precatalytic system was dissolved in<br />

[D 8]iPrOH and treated with [D 7]iPrONa for 5 min at 100 8C.<br />

After addition of 1, the solution was stirred for 5 h at<br />

100 8C. Only alcohol 17 was detected as product by 1 H NMR<br />

spectroscopy (Scheme 2; >99 %). [25] This result indicates an<br />

exclusive transfer of the deuterium into the carbonyl group.<br />

Apparently no C H activation processes and enol formation<br />

occurred under the described conditions. [26]<br />

To clarify the pathway of hydrogen transfer from the hydrogen<br />

donor to the substrate molecule, we used [D]iPrOH<br />

(the hydroxy group was deuterated) as solvent/donor and<br />

sodium 2-propylate as base in the transfer hydrogenation of<br />

1 (Scheme 2). We obtained a mixture of two different deuterated<br />

1-phenylethanols 18 and 19 in the ratio 85:15. [27]<br />

Chem. Asian J. 2006, 1, 598 – 604 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemasianj.org 601


FULL PAPERS<br />

Table 5. ACHTUNGTRENUNG[Fe 3(CO) 12]/terpy/PPh 3- and FeCl 2/terpy/PPh 3-catalyzed<br />

ACHTUNGTRENUNGhydrogenation of ketones. [a]<br />

Entry Alcohol Yield [%] [b]<br />

ACHTUNGTRENUNG[Fe 3(CO) 12]<br />

1 95 91<br />

2 > 99 97<br />

3 84 83<br />

4 63 75<br />

5 > 99 > 99<br />

6 81 92<br />

7 5 8<br />

8 48 57<br />

9 95 93<br />

10 > 99 > 99<br />

Yield [%] [b]<br />

FeCl 2<br />

[a] Reaction conditions: in situ catalyst (0.0038 mmol) ([Fe 3(CO) 12]<br />

(0.0013 mmol) or FeCl 2 (0.0038 mmol), terpy (0.0038 mmol), PPh 3<br />

(0.0038 mmol) in 2-propanol (2.0 mL), 16 h at 658C), iPrONa<br />

(0.38 mmol), 5 min at 100 8C, then addition of ketone (0.76 mmol), reaction<br />

for 7 h at 100 8C. [b] Yield was determined by GC (2: 50 m Lipodex<br />

E, 95–200 8C, 8: 25 m Lipodex E, 100 8C, 9: 50 m Lipodex E, 90–105 8C,<br />

10: 25 m Lipodex E, 80–180 8C, 11, 14, 15, and 16: 30 m, HP Agilent<br />

Technologies, 50–300 8C, 12: 25 m Lipodex E, 90–180 8C, 13: 50 m Lipodex<br />

E, 90–180 8C) with diglyme as internal standard (yield is equivalent<br />

to conversion).<br />

This specific migration is in agreement with the described<br />

monohydride mechanism, which implies that a major formation<br />

of the metal monohydride in the catalytic cycle occurred,<br />

albeit with a small amount of 19. [8c] This H/D scrambling<br />

is explained by the reversibility of the hydrogen-transfer<br />

process due to the hydrogen-donating ability of<br />

1-phenylethanol (low oxidation potential). [8c]<br />

Scheme 1. Monohydride (pathway A) and dihydride (pathway B) mechanisms<br />

of transition-metal-catalyzed transfer hydrogenation of acetophenone.<br />

Scheme 2. Deuterium incorporation into acetophenone (1) catalyzed by<br />

[Fe 3(CO) 12]/terpy/PPh 3 in the presence of base.<br />

Conclusions<br />

We have developed the first general homogeneous Fe catalyst<br />

system for the transfer hydrogenation of aliphatic and<br />

aromatic ketones. In the presence of 1 mol % of [Fe 3(CO) 12]/<br />

terpy/PPh 3 or FeCl 2/terpy/PPh 3, the corresponding alcohols<br />

are obtained in good to excellent yield and chemoselectivity.<br />

The active catalyst systems are easily generated in the presence<br />

of cheap available nitrogen and phosphorus ligands.<br />

Mechanistic experiments indicate a transfer of hydrogen<br />

from the donor molecule to the substrate by a monohydride<br />

mechanism. Further work in the direction of stereoselective<br />

Fe-based hydrogenation catalysts is under way in our laboratories.<br />

General<br />

Experimental Section<br />

M. Beller et al.<br />

All manipulations were performed under argon atmosphere with standard<br />

Schlenk techniques. Unless otherwise specified, all chemicals are<br />

commercially available and used as received. 2-Propanol, pyridine, and<br />

triethylamine were used without further purification (purchased from<br />

Fluka, dried over molecular sieves). Sodium 2-propylate and sodium tertbutylate<br />

were prepared by treating sodium with 2-propanol or tert-butanol<br />

under argon atmosphere (stock solution). Ketones 1, 8, 9, 11, 12, 14,<br />

602 www.chemasianj.org 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Asian J. 2006, 1, 598 – 604


Ketone Hydrogenation with Homogeneous Iron Catalysts<br />

15, and 16 were dried over CaH 2, distilled under vacuum, and stored<br />

under argon. Ketones 10 and 13 were treated with vacuum/argon cycles<br />

and stored under argon. Tmeda was distilled under Argon. [Et 3NH]-<br />

ACHTUNGTRENUNG[HFeCO 4], [13] BuP(adamantyl) 2, [28] and N-phenyl-2-(di-tert-butylphos-<br />

ACHTUNGTRENUNGphanyl)pyrrole [29] were synthesized according to literature protocols.<br />

[D 8]- and [D]iPrOH were dried over CaH 2 and distilled under argon atmosphere.<br />

General Procedure for Transfer Hydrogenation of Ketones<br />

In a Schlenk tube (10 mL), the catalyst (0.0038 mmol) was generated<br />

in situ by stirring a solution of [Fe 3(CO) 12] (0.0013 mmol), terpy<br />

(0.0038 mmol), and PPh 3 (0.0038 mmol) in 2-propanol (1.0 mL) for 16 h<br />

at 658C. The precatalytic system was treated with sodium 2-propylate<br />

(0.38 mmol) at 100 8C for 5 min. After addition of the corresponding<br />

ketone (0.38 or 0.76 mmol), the reaction mixture was stirred for 7 h at<br />

100 8C. The solution was cooled to room temperature and filtered over a<br />

plug of silica. The conversion was measured by GC without further purification.<br />

Procedures for Distinguishing Homogeneous and Heterogeneous Catalysts<br />

Mercury poisoning: In a Schlenk tube (10 mL), a solution of [Fe3(CO) 12]<br />

(0.0013 mmol), terpy (0.0038 mmol), and PPh3 (0.0038 mmol) in 2-propanol<br />

(1.0 mL) was stirred for 16 h at 658C. The precatalyst was treated at<br />

100 8C with sodium 2-propylate (0.38 mmol) in 2-propanol (0.5 mL) for<br />

5 min, followed by 1 (0.76 mmol) in 2-propanol (0.5 mL). The reaction<br />

mixture was kept for 1 h at 100 8C. After that, a drop of mercury, which<br />

was degassed and stored under argon, was added, and the reaction was<br />

continued for 7 h. The reaction mixture was cooled to room temperature<br />

and filtered over a plug of silica gel. The conversion was determined by<br />

GC without further purification.<br />

Maitlis test: In a Schlenk tube (10 mL), a solution of [Fe3(CO) 12]<br />

(0.0013 mmol), terpy (0.0038 mmol), and PPh3 (0.0038 mmol) in 2-propanol<br />

(1.0 mL) was stirred for 16 h at 658C. The precatalyst was treated at<br />

100 8C with sodium 2-propylate (0.38 mmol) in 2-propanol (0.5 mL) for<br />

5 min, followed by 1 (0.76 mmol) in 2-propanol (0.5 mL). The reaction<br />

mixture was kept for 1 h at 100 8C. After that, the solution was filtered<br />

under an atmosphere of argon through a filter supported by a plug of<br />

celite (heated in vacuum and stored under argon). The filtrate was<br />

heated again to 100 8C and the reaction allowed to proceed for another<br />

6 h. The celite phase was transferred into a Schlenk tube (10 mL) and<br />

mixed with 2-propanol (1.0 mL), sodium 2-propylate (0.38 mmol) in 2propanol<br />

(0.5 mL), and 1 (0.76 mmol) in 2-propanol (0.5 mL). After the<br />

reaction was complete, both mixtures were cooled to room temperature<br />

and filtered over a plug of silica gel. The conversion was determined by<br />

GC without further purification.<br />

Transfer Hydrogenation with [D 8]iPrOH as Hydride Source<br />

In a Schlenk tube (10 mL), [Fe 3(CO) 12] (0.0013 mmol), terpy<br />

(0.0038 mmol), and PPh 3 (0.0038 mmol) in [D 8]iPrOH (1.0 mL) were stirred<br />

for 16 h at 658C. After mixing the precatalytic solution with<br />

[D 7]iPrONa (0.38 mmol, prepared by reacting sodium (0.38 mmol) with<br />

[D 8]iPrOH (0.5 mL)) at 100 8C for 5 min, a solution of 1 (0.76 mmol) in<br />

[D 8]iPrOH (0.5 mL) was added. The reaction mixture was kept for 5 h at<br />

100 8C, then cooled to room temperature and filtered over a plug of<br />

silica. The conversion was determined by 1 H NMR spectroscopy.<br />

Transfer Hydrogenation with [D]iPrOH as Hydride Source<br />

In a Schlenk tube (10 mL), [Fe 3(CO) 12] (0.0013 mmol), terpy<br />

(0.0038 mmol), and PPh 3 (0.0038 mmol) in [D]iPrOH (1.0 mL) were stirred<br />

for 16 h at 658C. After mixing the precatalytic solution with sodium<br />

2-propylate (0.38 mmol, prepared by treating sodium (0.38 mmol) with<br />

[D]iPrOH (0.5 mL)) at 100 8C for 5 min, a solution of 1 (0.76 mmol) in<br />

[D]iPrOH (0.5 mL) was added. The reaction mixture was kept for 5 h at<br />

100 8C. (To avoid side effects, for instance, scrambling, the reaction was<br />

stopped before <strong>full</strong> conversion.) The solution was cooled to room temperature<br />

and filtered over a plug of silica. The solvent was removed under<br />

vacuum, and the residue was dissolved in CDCl 3. The conversion was de-<br />

termined by 1 H NMR spectroscopy. The ratio of 18 to 19 was based on<br />

the integrals of the 1 H NMR signals of the CH 3 groups.<br />

Acknowledgements<br />

This work was financed by the State of Mecklenburg-Vorpommern and<br />

the Bundesministerium für Bildung und Forschung (BMBF). We thank<br />

Mrs. C. Voss, Mrs. C. Mewes, Mrs. M. Heyken, Mrs. S. Buchholz, and Dr.<br />

C. Fischer (all Leibniz-Institut für Katalyse e.V. an der <strong>Universität</strong> <strong>Rostock</strong>)<br />

for their excellent technical and analytical support.<br />

[1] a) P. T. Anastas, M. M. Kirchhoff, Acc. Chem. Res. 2002, 35, 686 –<br />

694; b) P. T. Anastas, M. M. Kirchhoff, T. C. Williamson, Appl.<br />

Catal. A 2001, 221, 3–13.<br />

[2] The cost of iron is low relative to that of rhodium, ruthenium, or iridium<br />

(current price for 1 ton of iron US$300), due to the widespread<br />

abundance of iron in the earth’s crust and its easy accessibility.<br />

Furthermore, iron is an essential mineral for life (daily dose for<br />

humans 5–28 mg) and plays an important role in a large number of<br />

biological processes.<br />

[3] For an excellent review, see: C. Bolm, J. Legros, J. Le Paih, L. Zani,<br />

Chem. Rev. 2004, 104, 6217 –6254.<br />

[4] a) S. C. Bart, E. J. Hawrelak, E. Lobkovsky, P. J. Chirik, Organometallics<br />

2005, 24, 5518 –5527; b) S. C. Bart, E. Lobkovsky, P. J. Chirik,<br />

J. Am. Chem. Soc. 2004, 126, 13794 – 13807; c) Q. Knijnenburg,<br />

A. D. Horton, H. van der Heijden, W. A. Gal, P. Budzelaar, M. Henricus,<br />

PCT Int. Appl. WO 2003042131, 2003; d) H. Nagorski, M. J.<br />

Mirbach, J. Organomet. Chem. 1985, 291, 199 –204; e) M. E. Miller,<br />

E. R. Grant, J. Am. Chem. Soc. 1984, 106, 4635 – 4636; f) T. J. Lynch,<br />

M. Banah, H. D. Kaesz, C. R. Porter, J. Org. Chem. 1984, 49, 1266 –<br />

1270; g) M. A. Rafhi, L. Markó, J. Organomet. Chem. 1984, 262,<br />

359 – 364; h) H. Inoue, M. Sato, J. Chem. Soc. Chem. Commun. 1983,<br />

983 – 984; i) L. Markó, M. A. Radhi, I. Ötvçs, J. Organomet. Chem.<br />

1981, 218, 369 –376.<br />

[5] a) S. R. Boothroyd, M. A. Kerr, Tetrahedron Lett. 1995, 36, 2411 –<br />

2414; b) H. Alper, K. E. Hashem, J. Am. Chem. Soc. 1981, 103,<br />

6514 – 6515; c) J. F. Knifton, J. Org. Chem. 1976, 41, 1200 –1206.<br />

[6] I. Iovel, K. Mertins, J. Kischel, A. Zapf, M. Beller, Angew. Chem.<br />

2005, 117, 3981 –3985; Angew. Chem. Int. Ed. 2005, 44, 3913 –3916.<br />

[7] a) K. Junge, G. Oehme, A. Monsees, T. Riermeier, U. Dingerdissen,<br />

M. Beller, Tetrahedron Lett. 2002, 43, 4977 –4980; b) K. Junge, G.<br />

Oehme, A. Monsees, T. Riermeier, U. Dingerdissen, M. Beller, J.<br />

Organomet. Chem. 2003, 675, 91–96; c) K. Junge, B. Hagemann, S.<br />

Enthaler, A. Spannenberg, M. Michalik, G. Oehme, A. Monsees, T.<br />

Riermeier, M. Beller, Tetrahedron: Asymmetry 2004, 15, 2621 –2631;<br />

d) K. Junge, B. Hagemann, S. Enthaler, G. Oehme, M. Michalik, A.<br />

Monsees, T. Riermeier, U. Dingerdissen, M. Beller, Angew. Chem.<br />

2004, 116, 5176 – 5179; Angew. Chem. Int. Ed. 2004, 43, 5066 –5069;<br />

e) B. Hagemann, K. Junge, S. Enthaler, M. Michalik, T. Riermeier,<br />

A. Monsees, M. Beller, Adv. Synth. Catal. 2005, 347, 1978 –1986;<br />

f) S. Enthaler, B. Hagemann, K. Junge, G. Erre, M. Beller, Eur. J.<br />

Org. Chem. 2006, 2912 –2917.<br />

[8] a) G. Zassinovich, G. Mestroni, S. Gladiali, Chem. Rev. 1992, 92,<br />

1051 – 1069; b) R. Noyori, S. Hashiguchi, Acc. Chem. Res. 1997, 30,<br />

97–102; c) S. Gladiali, G. Mestroni in Transition Metals for Organic<br />

Synthesis, 2nd ed. (Eds.: M. Beller, C. Bolm), Wiley-VCH, Weinheim,<br />

2004, pp. 145 –166; d) H.-U. Blaser, C. Malan, B. Pugin, F.<br />

Spindler, M. Studer, Adv. Synth. Catal. 2003, 345, 103 –151; e) S.<br />

Gladiali, E. Alberico, Chem. Soc. Rev. 2006, 35, 226 –236.<br />

[9] R. Noyori, I. Umeda, T. Ishigami, J. Org. Chem. 1972, 37, 1542 –<br />

1545.<br />

[10] a) K. Jothimony, S. Vancheesan, J. Mol. Catal. 1989, 52, 301 – 304;<br />

b) K. Jothimony, S. Vancheesan, J. C. Kuriacose, J. Mol. Catal. 1985,<br />

32, 11 –16.<br />

[11] J.-S. Chen, L.-L. Chen, Y. Xing, G. Chen, W.-Y. Shen, Z.-R. Dong,<br />

Y.-Y. Li, J.-X. Gao, Acta Chimica Sinica, 2004, 62, 1745 –1750.<br />

Chem. Asian J. 2006, 1, 598 – 604 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemasianj.org 603


FULL PAPERS<br />

[12] C. Bianchini, E. Farnetti, M. Graziani, M. Peruzzini, A. Polo, Organometallics<br />

1993, 12, 3753 – 3761.<br />

[13] L. F. Dahl, J. F. Blount, Inorg. Chem. 1965, 4, 1373 – 1375.<br />

[14] H. Zhang, C.-B. Yang, Y.-Y. Li, Z.-R. Donga, J.-X. Gao, H. Nakamura,<br />

K. Murata, T. Ikariya, Chem. Commun. 2003, 142 –143.<br />

[15] a) C. Milone, R. Ingoglia, L. Schipilliti, C. Crisa<strong>full</strong>i, G. Neri, S. Galvagno,<br />

J. Catal. 2005, 236, 80–90; b) C. Milone, R. Ingoglia, M. L.<br />

Tropeano, G. Neri, S. Galvagno, Chem. Commun. 2003, 868 – 869;<br />

c) S. Narayanan, R. Unnikrishnan, J. Chem. Soc. Faraday Trans.<br />

1998, 94, 1123 –1128; d) H. Itoh, H. Hosaka, T. Ono, E. Kikuchi,<br />

Appl. Catal. 1988, 40, 53–66; < lit e > L. S. Glebov, A. I. Mikaya,<br />

V. I. Smetanin, V. G. Zaikin, G. A. Kliger, S. M. Loktev, J. Catal.<br />

1985, 93, 75–82.<br />

[16] G. M. Whitesides, M. Hackett, R. L. Brainard, J. P. P. M. Lavalleye,<br />

A. F. Sowinski, A. N. Izumi, S. S. Moore, D. W. Brown, E. M. Staudt,<br />

Organometallics 1985, 4, 1819 – 1830.<br />

[17] a) C. M. Hagen, L. Vieille-Petit, G. Laurenczy, G. Süss-Fink, R. G.<br />

Finke, Organometallics 2005, 24, 1819 – 1831; b) J. A. Widegren,<br />

R. G. Finke, J. Mol. Catal. A: Chem. 2003, 198, 317 –341.<br />

[18] J. E. Hamlin, K. Hirai, A. Millan, P. M. Maitlis, J. Mol. Catal. 1980,<br />

7, 543 –544.<br />

[19] The preparation of the precatalyst was carried out in a similar<br />

manner to the described procedure, with [Fe 3(CO) 12], terpy, and<br />

PPh 3, stirring for 16 h at 658C in[D 8]iPrOH, addition of sodium 2propylate<br />

(5 equiv), and transfer into an NMR tube. The suspension<br />

was heated to 100 8C, and after 5 min the reaction mixture was<br />

quickly cooled to room temperature and immediately subjected to<br />

1 Hor 31 P NMR spectroscopy.<br />

[20] The preparation of the precatalyst was carried out in a manner similar<br />

to the described procedure, with [Fe 3(CO) 12], terpy, and PPh 3,<br />

stirring for 16 h at 65 8C in[D 8]iPrOH or 2-propanol, followed by<br />

transfer of the suspension into an NMR tube or removal of 2-propanol<br />

and dissolution of the residue in [D 4]MeOH, CDCl 3,<br />

[D 6]acetone, or [D 6]benzene.<br />

[21] a) D. D. Tanner, G. E. Diaz, A. Potter, J. Org. Chem. 1985, 50,<br />

2149 – 2154; b) M. Degueil-Castaing, A. Rahm, J. Org. Chem. 1986,<br />

51, 1672 –1676;c) T. Ohkuma, M. Koizumi, H. Doucet, T. Pham, M.<br />

Kozawa, K. Murata, E. Katayama, T. Yokozawa, T. Ikariya, R.<br />

Noyori, J. Am. Chem. Soc. 1998, 120, 13 529 –13 530; d) J. Wu, J.-X.<br />

Ji, R. Guo, C.-H. Yeung, A. S. C. Chan, Chem. Eur. J. 2003, 9, 2963 –<br />

2968.<br />

[22] E. C. Ashby, J. N. Argyropoulos, J. Org. Chem. 1986, 51, 3593 –3597.<br />

[23] a) For recent reviews, see: J. S. M. Samec, J.-E. Bäckvall, P. G. Andersson,<br />

P. Brandt, Chem. Soc. Rev. 2006, 35, 237 – 248; b) T. Ikariya,<br />

K. Murata, R. Noyori, Org. Biomol. Chem. 2006, 4, 393 –406 and<br />

references therein; c) H. Guan, M. Iimura, M. P. Magee, J. R.<br />

M. Beller et al.<br />

Norton, G. Zhu, J. Am. Chem. Soc. 2005, 127, 7805 –7814; d) M.<br />

Gómez, S. Janset, G. Muller, G. Aullón, M. A. Maestro, Eur. J.<br />

Inorg. Chem. 2005, 4341 – 4351; e) K. MuÇiz, Angew. Chem. 2005,<br />

117, 6780 –6785; Angew. Chem. Int. Ed. 2005, 44, 6622 – 6627; f) C.<br />

Hedberg, K. Källstrçm, P. I. Arvidsson, P. Brandt, P. G. Andersson,<br />

J. Am. Chem. Soc. 2005, 127, 15083 –15090; g) A. S. Y. Yim, M.<br />

Wills, Tetrahedron 2005, 61, 7994 – 8004; h) S. E. Clapham, A. Hadzovic,<br />

R. H. Morris, Coord. Chem. Rev. 2004, 248, 2201 –2237; i) T.<br />

Koike, T. Ikariya, Adv. Synth. Catal. 2004, 346, 37 –41; j) P. Brandt,<br />

P. Roth, P. G. Andersson, J. Org. Chem. 2004, 69, 4885 –4890; k) J.-<br />

W. Handgraaf, J. N. H. Reek, E. J. Meijer, Organometallics 2003, 22,<br />

3150 – 3157; l) C. P. Casey, J. B. Johnson, J. Org. Chem. 2003, 68,<br />

1998 – 2001; m) C. A. Sandoval, T. Ohkuma, K. MuÇiz, R. Noyori, J.<br />

Am. Chem. Soc. 2003, 125, 13490 –13503; n) R. Noyori, Angew.<br />

Chem. 2002, 114, 2108 – 2123; Angew. Chem. Int. Ed. 2002, 41, 2008 –<br />

2022; o) C. P. Casey, S. W. Singer, D. R. Powell, R. K. Hayashi, M.<br />

Kavana, J. Am. Chem. Soc. 2001, 123, 1090 –1100; p) O. Pàmies,<br />

J. E. Bäckvall, Chem. Eur. J. 2001, 7, 5052 –5058; q) C. S. Yi, Z. He,<br />

Organometallics 2001, 20, 3641 –3643; r) R. Noyori, M. Yamakawa,<br />

S. Hashiguchi, J. Org. Chem. 2001, 66, 7931 – 7944; s) M. Yamakawa,<br />

H. Ito, R. Noyori, J. Am. Chem. Soc. 2000, 122, 1466 –1478; t) D. A.<br />

Alonso, P. Brandt, S. J. M. Nordin, P. G. Andersson, J. Am. Chem.<br />

Soc. 1999, 121, 9580 –9588; u) D. G. I. Petra, J. N. H. Reek, J.-W.<br />

Handgraaf, E. J. Meijer, P. Dierkes, P. C. J. Kamer, J. Brussee, H. E.<br />

Schoemaker, P. W. N. M. van Leeuwen, Chem. Eur. J. 2000, 6, 2818 –<br />

2829; v) A. Aranyos, G. Csjernyik, K. J. Szabó, J. E. Bäckvall, Chem.<br />

Commun. 1999, 351 –352; w) M. L. S. Almeida, M. Beller, G. Z.<br />

Wang, J. E. Bäckvall, Chem. Eur. J. 1996, 2, 1533 –1536.<br />

[24] a) L. Dahlenburg, R. Gçtz, Eur. J. Inorg. Chem. 2004, 88 –905;<br />

b) P. W. C. Cross, G. J. Ellames, J. S. Gibson, J. M. Herbert, W. J.<br />

Kerr, A. H. McNeill, T. W. Mathers, Tetrahedron 2003, 59, 3349 –<br />

3358.<br />

[25] a) L. Y. Kuo, D. M. Finigan, N. N. Tadros, Organometallics 2003, 22,<br />

2422 – 2425; b) R. L. Elsenbaumer, H. S. Mosher, J. Org. Chem.<br />

1979, 44, 600 –604.<br />

[26] D. Klomp, T. Maschmeyer, U. Hanefeld, J. A. Peters, Chem. Eur. J.<br />

2004, 10, 2088 –2093.<br />

[27] The ratio of 18 to 19 was determined by 1 H NMR spectroscopy<br />

based on the integrals of the signals for the CH 3 groups.<br />

[28] A. Ehrentraut, A. Zapf, M. Beller, Synlett 2000, 1589 – 1592.<br />

[29] A. Zapf, R. Jackstell, F. Rataboul, T. Riermeier, A. Monsees, C.<br />

Fuhrmann, N. Shaikh, U. Dingerdissen, M. Beller, Chem. Commun.<br />

2004, 38 –39.<br />

Received: April 4, 2006<br />

Revised: July 5, 2006<br />

604 www.chemasianj.org 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Asian J. 2006, 1, 598 – 604


6. Transfer hydrogenation<br />

6.1. Homogeneous iron catalyst in reduction chemistry<br />

6.1.2. Biomimetic transfer hydrogenation of ketones with iron<br />

porphyrin catalysts<br />

Stephan Enthaler, Giulia Erre, Man Kin Tse, Kathrin Junge, and Matthias Beller*,<br />

Tetrahedron Lett. 2006, 47, 8095–8099.<br />

Contributions<br />

Ligands 1a and 1d were synthesized by Dr. M. K. Tse.<br />

108


Biomimetic transfer hydrogenation of ketones with iron<br />

porphyrin catalysts<br />

Stephan Enthaler, Giulia Erre, Man Kin Tse, Kathrin Junge and Matthias Beller *<br />

Leibniz-Institut für Katalyse e.V. an der <strong>Universität</strong> <strong>Rostock</strong>, Albert-Einstein-Str. 29a, D-18059 <strong>Rostock</strong>, Germany<br />

Received 17 July 2006; revised 11 September 2006; accepted 12 September 2006<br />

Available online 4 October 2006<br />

Abstract—For the first time in situ generated iron porphyrins have been applied as homogeneous catalysts for the transfer hydrogenation<br />

of ketones. Using 2-propanol as hydrogen source various ketones are reduced to the corresponding alcohols in good to<br />

excellent yield and selectivity. Under optimized reaction conditions high catalyst turnover frequencies up to 642 h 1 are achieved.<br />

Ó 2006 Elsevier Ltd. All rights reserved.<br />

Alcohols are key intermediates for the synthesis of pharmaceuticals,<br />

agrochemicals, polymers and new materials.<br />

1,2 Starting from carbonyl compounds various<br />

catalytic approaches toward the synthesis of alcohols<br />

have been developed. Typical examples are the addition<br />

of organometallic compounds to aldehydes, hydrosilylation,<br />

and hydrogenation of aldehydes or ketones. 2<br />

Among these transformations hydrogenations represent<br />

the most atom-efficient and environmentally benign<br />

methodology. In particular, transfer hydrogenation is<br />

a powerful strategy because of the ease of performance<br />

and general applicability. 3 More specifically, a broad<br />

scope of alcohols is available by transfer hydrogenation<br />

using non-toxic hydrogen donors, for example, 2-propanol<br />

or HCOOH/NEt3, under mild reaction conditions in<br />

the presence of precious metal catalysts based on Ir, Rh,<br />

Ru, or Ni. 4<br />

With regard to the upcoming catalyst developments a<br />

fundamental challenge is the substitution of these expensive<br />

and rare transition metals by less toxic, inexpensive<br />

and abundantly available metals such as iron. 5 Until<br />

now, homogeneous iron catalysts have been most frequently<br />

applied for carbon–carbon coupling reactions,<br />

such as olefin polymerizations, cross-couplings and cycloadditions<br />

as well as reductions of nitro compounds. 6<br />

However, much less attention was directed toward<br />

iron-catalyzed (transfer) hydrogenations, although the<br />

relevance of such reductions is evident even with respect<br />

* Corresponding author. Tel.: +49 381 1281 113; fax: +49 381 1281<br />

5000; e-mail: matthias.beller@catalysis.de<br />

0040-4039/$ - see front matter Ó 2006 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.tetlet.2006.09.058<br />

Tetrahedron Letters 47 (2006) 8095–8099<br />

to industrial applications. To the best of our knowledge<br />

only the groups of Noyori, 7 Vancheesan, 8 Bianchini 9<br />

and Gao 10 reported the utilization of iron salts and iron<br />

complexes in the reduction of a,b-unsaturated carbonyl<br />

compounds and ketones. For tuning the activity and<br />

selectivity of these catalysts, oxygen and moisture sensitive<br />

tetradentate phosphines 9 or aminophosphines 10<br />

have been predominantly applied as ligands.<br />

Inspired by nature we thought that multidentate nitrogen<br />

ligands should be also suitable ligands for stabilizing<br />

iron as metal centre in transfer hydrogenations. In the<br />

past biomimetic ligand toolboxes were invented, which<br />

are based on natural sources or similar structural motifs,<br />

for example, amino alcohols, quinidines, bisoxazolines<br />

and porphyrins. 11 Within these potential ligands,<br />

porphyrins, which are involved in manifold biological<br />

redox processes, seemed of special interest to us due to<br />

their strong ability to stabilize the iron centre. 12,13<br />

Stimulated by our ongoing research in catalytic hydrogenations<br />

15 we became interested in developing new<br />

hydrogenation catalysts based on iron. Thus, we report<br />

herein for the first time a combination of iron and porphyrins<br />

as catalysts for the efficient reduction of various<br />

ketones (Scheme 1).<br />

In exploratory experiments, 2-propanol-based transfer<br />

hydrogenation of acetophenone was examined using<br />

an easy to adopt in situ catalyst system comprising<br />

Fe 3(CO) 12 and porphyrin 1a. Typically, the active catalyst<br />

is prepared by stirring a solution of 1 mol % iron<br />

source and 1 mol % 1a in 2-propanol (1.0 mL) for 16 h


8096 S. Enthaler et al. / Tetrahedron Letters 47 (2006) 8095–8099<br />

Scheme 1. Selection of applied porphyrins. 14<br />

at 65 °C. After the addition of 50 mol % base the<br />

mixture is heated for 5 min at 100 °C and the standard<br />

substrate acetophenone 4 is added. 16<br />

Initially, the influence of temperature and various bases<br />

on the reaction rate was investigated. An optimal catalyst<br />

activity is obtained at 100 °C (Table 1, entries 1–<br />

3). In general, sodium and potassium alkoxides gave<br />

an excellent yield of 1-phenylethanol (96–99%, Table<br />

1, entries 4–8). Noteworthy from a practical point of<br />

view is that NaOH and K 2CO 3 also led to a high product<br />

yield. However, in the presence of organic bases such<br />

as NEt3 and pyridine (Table 1, entries 10 and 11) no<br />

significant amount of product is formed in reasonable<br />

time. Diminishing the base concentration resulted in a<br />

decrease of 1-phenylethanol. In the absence of a base,<br />

no transfer of hydrogen was observed.<br />

Table 1. Catalytic transfer hydrogenation of acetophenone 4<br />

Entry Iron source Base Temperature (°C) Yield (%) a<br />

1 Fe3(CO)12 2-PrONa 80 56<br />

2 Fe3(CO) 12 2-PrONa 90 68<br />

3 Fe3(CO) 12 2-PrONa 100 96<br />

4 Fe3(CO)12 NaOH 100 98<br />

5 Fe3(CO)12 KOH 100 42<br />

6 Fe3(CO) 12 LiOH 100 5<br />

7 Fe3(CO) 12 K-t-OBu 100 99<br />

8 Fe3(CO)12 Na-t-OBu 100 97<br />

9 Fe3(CO)12 K2CO3 100 89<br />

10 Fe3(CO) 12 NEt3 100 1%) of by-products (e.g., aldol condensation<br />

products) are obtained.<br />

Next, attempts were made concerning the iron source<br />

(Table 2). 17 The best activity was obtained for<br />

Fe 3(CO) 12, FeBr 2, Fe(acac) 3 and [Et 3NH][HFe(CO) 4]. 18<br />

Surprisingly, no reliable correlation between the oxidation<br />

state and reaction rate was observed as high activity<br />

was detected for Fe(0)-, Fe(II)- and Fe(III)-salts. Hence,<br />

the formation of the active catalyst species is complete<br />

for the various pre-catalysts under the applied conditions.<br />

However, studying the dependency of conversion<br />

versus reaction time in the presence of Fe3(CO)12<br />

revealed an induction period of nearly 2 h.<br />

Table 2. Influence of iron sources in the transfer hydrogenation of<br />

acetophenone 4<br />

Entry Iron source Yield (%) a<br />

1 Fe3(CO) 12 96<br />

2 FeBr2 98<br />

3 FeCl2 90<br />

4 FeCl3 86<br />

5 Fe(acac) 2 74<br />

6 Fe(acac)3 97<br />

7 CpFe(CO)2I 44<br />

8 FeSO4 46<br />

9 [Et3NH][HFe(CO) 4] 97<br />

Standard reaction conditions: 0.0038 mmol in situ catalyst<br />

(0.0038 mmol Fe-source or 0.0013 mmol Fe3(CO)12 and 0.0038 mmol<br />

1a in 2.0 mL 2-propanol for 16 h at 65 °C), 0.19 mmol base, 5 min at<br />

100 °C, then the addition of 0.38 mmol acetophenone 4, 7 h at 100 °C.<br />

a<br />

Yield and conversion were determined by GC analysis (50 m Lipodex<br />

E, 95–150 °C) with diglyme as an internal standard.


In the case of Fe(acac)2 and Fe(acac)3 a favourable<br />

conversion was observed for the Fe(III)-salt. Comparing<br />

the results of FeCl 2 and FeBr 2, a small influence of the<br />

corresponding halide was also detected.<br />

Next, we focused our attention on the influence of the<br />

metal–ligand ratio. Even without any ligand some catalytic<br />

activity is obtained (54% of 5). The addition of 0.5<br />

or 0.2 equiv of 1a (with respect to Fe) increased the yield<br />

of 1-phenylethanol to 74% and 86%, respectively. The<br />

highest yield is obtained at a metal–ligand ratio of 1:1.<br />

A further increase of the number of ligand with respect<br />

to iron (2:1) resulted in a slight decrease of activity (89%<br />

of 5).<br />

In order to further improve the catalyst system, we<br />

applied different porphyrin ligands (1a–3) in the model<br />

reaction in the presence of 0.5 mol % Fe catalyst<br />

(Fe 3(CO) 12, sodium 2-propylate, 100 °C). As shown in<br />

Table 3, best yields were achieved with meso-substituted<br />

porphyrins 1b and 1e (Table 3, entries 3 and 6). Substitution<br />

in the meso-phenylic system of porphyrin 1b with<br />

electron withdrawing groups (1a, 1c, and 1d) displayed a<br />

decrease in activity (Table 3, entries 1, 3 and 4).<br />

In addition, porphyrins substituted in the b-pyrrolenic<br />

positions were employed in the reduction of acetophenone.<br />

The activity of the symmetric coproporphyrin I<br />

2 was to some extend lower when compared with<br />

meso-substituted porphyrins. The natural complex chloroprotoporphyrin<br />

IX Fe(III) 3 was utilized without catalysts<br />

pre-formation as described for all other ligands,<br />

and a moderate yield of 1-phenylethanol was detected<br />

(Table 3, entry 8). Nevertheless, complex 3 is an interesting<br />

catalyst for such reactions due to easier handling and<br />

availability.<br />

Table 3. Testing of different porphyrins in the transfer hydrogenation<br />

of acetophenone 4<br />

Entry Porphyrin Catalyst<br />

loading (mol %)<br />

Yield (%) a TOF (h 1 ) b<br />

1 1a 0.5 90 26<br />

2 1a 0.01 45 642<br />

3 1b 0.5 93 27<br />

4 1c 0.5 68 19<br />

5 1d 0.5 56 16<br />

6 1e 0.5 94 27<br />

7 2 0.5 45 13<br />

8 3 0.5 51 15<br />

Standard reaction conditions: 0.0038 mmol or 0.000038 mmol in situ<br />

catalyst (0.0013 mmol or 0.000013 mmol Fe3(CO)12 and 0.0038 mmol<br />

or 0.000038 mmol porphyrin in 2.0 mL 2-propanol for 16 h at 65 °C),<br />

0.19 mmol or 0.0019 mmol sodium 2-propylate, 5 min at 100 °C, then<br />

the addition of 0.76 mmol acetophenone 4, 7 h at 100 °C.<br />

a<br />

Conversion was determined by GC analysis (50 m Lipodex E, 95–<br />

150 °C) with diglyme as an internal standard.<br />

b<br />

Turnover frequencies were determined after 7 h.<br />

S. Enthaler et al. / Tetrahedron Letters 47 (2006) 8095–8099 8097<br />

Table 4. Fe-catalyzed reduction of various ketones in the presence of<br />

porphyrins 1b and 1e<br />

Entry Product 1b Yield (%) a<br />

1 46 72<br />

2 >99 >99<br />

3 93 95<br />

4 50 68<br />

5 92 87<br />

6 21 22<br />

7


8098 S. Enthaler et al. / Tetrahedron Letters 47 (2006) 8095–8099<br />

6 and 7). Comparing different substitutions on the<br />

phenyl ring no reliable relationship between electron<br />

donating and electron withdrawing substituents and<br />

activity was observed (Table 4, entries 1–4). Similar to<br />

the model reaction, in all cases conversion and yield<br />

were nearly identical.<br />

In agreement with previous findings the highest yield is<br />

obtained with 2-methoxyacetophenone due to the presence<br />

of a second coordination site. 19<br />

In addition to aryl alkyl ketones, we also examined more<br />

challenging dialkyl ketones in this iron-catalyzed transfer<br />

hydrogenation. Good conversion and yield (89–<br />

90%) were observed for both substrates applying an iron<br />

catalyst containing 1e as ligand. In general, ligand 1e<br />

gave better results compared to 1b. This effect is especially<br />

pronounced for the dialkyl substrates (Table 4,<br />

entries 8 and 9).<br />

In conclusion, we have demonstrated for the first time<br />

the successful application of in situ prepared iron<br />

porphyrin catalysts in the transfer hydrogenation of<br />

ketones. The catalyst system is easily prepared and mimics<br />

biologically occurring Fe complexes. Under optimized<br />

conditions turnover frequencies up to 642 h 1<br />

were achieved. The scope and limitation of the catalyst<br />

were demonstrated on the reduction of nine different<br />

ketones with good to excellent yields.<br />

Acknowledgements<br />

This work has been financed by the State of Mecklenburg-Western<br />

Pomerania, the Bundesministerium für<br />

Bildung und Forschung (BMBF) and the Deutsche<br />

Forschungsgemeinschaft (Leibniz-award). We thank<br />

Mr. B. Hagemann, Mrs. C. Mewes, Mrs. M. Heyken,<br />

Mrs. S. Buchholz, and Dr. C. Fischer (all Leibniz-Institut<br />

für Katalyse e.V. an der <strong>Universität</strong> <strong>Rostock</strong>) for<br />

their excellent technical and analytical support.<br />

References and notes<br />

1. (a) Lennon, I. C.; Ramsden, J. A. Org. Process Res. Dev.<br />

2005, 9, 110–112; (b) Hawkins, J. M.; Watson, T. J. N.<br />

Angew. Chem. 2004, 116, 3286–3290; (c) Noyori, R.;<br />

Ohkuma, T. Angew. Chem. 2001, 113, 40–75; (d) Miyagi,<br />

M.; Takehara, J.; Collet, S.; Okano, K. Org. Process Res.<br />

Dev. 2000, 4, 346–348.<br />

2. (a) Transition Metals for Organic Synthesis, 2nd ed.;<br />

Beller, M., Bolm, C., Eds.; Wiley-VCH: Weinheim, 2004;<br />

(b) Cornils, B.; Herrmann, W. A. Applied Homogeneous<br />

Catalysis with Organometallic Compounds; Wiley-VCH:<br />

Weinheim, 1996; (c) Kitamura, M.; Noyori, R. In Ruthenium<br />

in Organic Synthesis; Murahashi, S.-I., Ed.; Wiley-<br />

VCH: Weinheim, 2004; (d) Comprehensive Asymmetric<br />

Catalysis; Jacobsen, E. N., Pfaltz, A., Yamamoto, H.,<br />

Eds.; Springer: Berlin, 1999; (e) Noyori, R. Asymmetric<br />

Catalysis in Organic Synthesis; Wiley: New York, 1994.<br />

3. (a) Zassinovich, G.; Mestroni, G.; Gladiali, S. Chem. Rev.<br />

1992, 51, 1051–1069; (b) Noyori, R.; Hashiguchi, S. Acc.<br />

Chem. Res. 1997, 30, 97–102; (c) Gladiali, S.; Mestroni, G.<br />

In Transition Metals for Organic Synthesis, 2nd ed.; Beller,<br />

M., Bolm, C., Eds.; Wiley-VCH: Weinheim, 2004; pp 145–<br />

166; (d) Blaser, H.-U.; Malan, C.; Pugin, B.; Spindler, F.;<br />

Studer, M. Adv. Synth. Catal. 2003, 345, 103–151; (e)<br />

Gladiali, S.; Alberico, E. Chem. Soc. Rev. 2006, 35, 226–<br />

236; (f) Samec, J. S. M.; Bäckvall, J.-E.; Andersson, P. G.;<br />

Brandt, P. Chem. Soc. Rev. 2006, 35, 237–248.<br />

4. Selected recent examples of transfer hydrogenations: (a)<br />

Schlatter, A.; Kundu, M. K.; Woggon, W.-D. Angew.<br />

Chem., Int. Ed. 2004, 43, 6731–6734; (b) Xue, D.; Chen,<br />

Y.-C.; Cui, X.; Wang, Q.-W.; Zhu, J.; Deng, J.-G. J. Org.<br />

Chem. 2005, 70, 3584–3591; (c) Wu, X.; Li, X.; King, F.;<br />

Xiao, J. Angew. Chem., Int. Ed. 2005, 44, 3407–3411; (d)<br />

Hayes, A. M.; Morris, D. J.; Clarkson, G. J.; Wills, M. J.<br />

Am. Chem. Soc. 2005, 127, 7318–7319; (e) Baratta, W.;<br />

Chelucci, G.; Gladiali, S.; Siega, K.; Toniutti, M.; Zanette,<br />

M.; Zangrando, E.; Rigo, P. Angew. Chem., Int. Ed. 2005,<br />

44, 6214–6219.<br />

5. Due to the widespread abundance of iron in earth’s crust<br />

and the easy accessibility the current price of iron is<br />

relatively low (approximately 300 US-dollar/t) compared<br />

to rhodium, iridium or ruthenium. Furthermore, iron is an<br />

essential trace element (daily dose for humans 5–28 mg)<br />

and is involved in an extensive number of biological<br />

processes.<br />

6. Bolm, C.; Legros, J.; Le Paih, J.; Zani, L. Chem. Rev.<br />

2004, 104, 6217–6254.<br />

7. Noyori, R.; Umeda, I.; Ishigami, T. J. Org. Chem. 1972,<br />

37, 1542–1545.<br />

8. (a) Jothimony, K.; Vancheesan, S. J. Mol. Catal. 1989, 52,<br />

301–304; (b) Jothimony, K.; Vancheesan, S.; Kuriacose, J.<br />

C. J. Mol. Catal. 1985, 32, 11–16.<br />

9. Bianchini, C.; Farnetti, E.; Graziani, M.; Peruzzini, M.;<br />

Polo, A. Organometallics 1993, 12, 3753–3761.<br />

10. Chen, J.-S.; Chen, L.-L.; Xing, Y.; Chen, G.; Shen, W.-Y.;<br />

Dong, Z.-R.; Li, Y.-Y.; Gao, J.-X. Huaxue Xuebao 2004,<br />

62, 1745–1750.<br />

11. (a) Jacqueline, H.-P. Chiral Auxiliaries and Ligands in<br />

Asymmetric Synthesis; John Wiley and Sons: New York,<br />

1995; (b) Larrow, J. F.; Jacobsen, E. N. Topics Organomet.<br />

Chem. 2004, 6, 123–152.<br />

12. (a) Porphyrins and Metalloporphyrins; Smith, K. M., Ed.;<br />

Elsevier: Netherlands, 1975; (b) The Porphyrins; Dolphin,<br />

D., Ed.; Academic Press: New York, 1978; (c) Che, C.-M.;<br />

Huang, J.-S. Coord. Chem. Rev. 2002, 231, 151–164; (d)<br />

Simonneaux, G.; Le Maux, P. Coord. Chem. Rev. 2002,<br />

228, 43–60; (e) Rose, E.; Andrioletti, B.; Zrig, S.; Quelquejeu-Etheve,<br />

M. Chem. Soc. Rev. 2005, 34, 573–583.<br />

13. (a) Munire, B. Chem. Rev. 1992, 92, 1411–1456; (b)<br />

Mansuy, D. Coor. Chem. Rev. 1993, 125, 129–141; (c)<br />

Metalloporphyrins in Catalytic Oxidations; Sheldon, R. A.,<br />

Ed.; M. Dekker Inc.: New York, 1994; (d) Che, C.-M.;<br />

Yu, W.-Y. Pure Appl. Chem. 1999, 71, 281–288; (e)<br />

Vinhado, F. S.; Martins, P. R.; Iamamoto, Y. Curr.<br />

Topics. Catal. 2002, 3, 199–213; (f) Rose, E.; Andrioletti,<br />

B.; Zrig, S.; Quelquejeu-Etheve, M. Chem. Soc. Rev. 2005,<br />

34, 573–583; (g) Shitama, H.; Katsuki, T. Tetrahedron<br />

Lett. 2006, 47, 3203–3207; (h) Zhou, C. Y.; Chan, P. W.<br />

H.; Che, C. M. Org. Lett. 2006, 8, 325–328; (i) Berkessel,<br />

A. Pure Appl. Chem. 2005, 77, 1277–1284.<br />

14. All porphyrins are commercial available by Strem or<br />

Aldrich except porphyrin 1d, which was synthesized<br />

according to a literature protocol: (a) Lindsey, J. S.;<br />

Hsu, H. C.; Schreiman, I. C. Tetrahedron Lett. 1986, 27,<br />

4969–4970; (b) Lindsey, J. S.; Schreiman, I. C.; Hsu, H. C.;<br />

Kearney, P. C.; Marguerettaz, A. M. J. Org. Chem. 1987,<br />

52, 827–836.<br />

15. (a) Junge, K.; Oehme, G.; Monsees, A.; Riermeier, T.;<br />

Dingerdissen, U.; Beller, M. Tetrahedron Lett. 2002, 43,


4977–4980; (b) Junge, K.; Oehme, G.; Monsees, A.;<br />

Riermeier, T.; Dingerdissen, U.; Beller, M. J. Organomet.<br />

Chem. 2003, 675, 91–96; (c) Junge, K.; Hagemann, B.;<br />

Enthaler, S.; Spannenberg, A.; Michalik, M.; Oehme, G.;<br />

Monsees, A.; Riermeier, T.; Beller, M. Tetrahedron:<br />

Asymmetry 2004, 15, 2621–2631; (d) Junge, K.; Hagemann,<br />

B.; Enthaler, S.; Oehme, G.; Michalik, M.;<br />

Monsees, A.; Riermeier, T.; Dingerdissen, U.; Beller, M.<br />

Angew. Chem. 2004, 116, 5176–5179; Angew. Chem., Int.<br />

Ed. 2004, 43, 5066–5069; (e) Hagemann, B.; Junge, K.;<br />

Enthaler, S.; Michalik, M.; Riermeier, T.; Monsees, A.;<br />

Beller, M. Adv. Synth. Catal. 2005, 347, 1978–1986; (f)<br />

Enthaler, S.; Hagemann, B.; Junge, K.; Erre, G.; Beller,<br />

M. Eur. J. Org. Chem. 2006, 2912–2917.<br />

16. All experiments were carried out under an inert gas<br />

atmosphere (argon) with exclusion of air. For the standard<br />

S. Enthaler et al. / Tetrahedron Letters 47 (2006) 8095–8099 8099<br />

reaction procedure Fe3(CO)12 (0.0013 mmol) and porphyrin<br />

1a (0.0038 mmol) is dissolved in 1.0 mL 2-propanol<br />

and stirred for 16 h at 65 °C. Then a solution of sodium 2propylate<br />

(0.19 mmol) in 0.5 mL 2-propanol is added. The<br />

solution is stirred for 5 min at 100 °C followed by the<br />

addition of 0.38 mmol acetophenone 4. After 7 h at<br />

100 °C, the mixture is cooled to rt and filtered over a plug<br />

of silica gel. The conversion and yield were determined<br />

by GC without further purification. All synthesized<br />

alcohols are known compounds. Their characterization<br />

is done by a comparison with authentic samples.<br />

17. Sodium 2-propylate was chosen due to easier handling.<br />

18. Dahl, L. F.; Blount, J. F. Inorg. Chem. 1965, 4, 1373–<br />

1375.<br />

19. Enthaler, S.; Hagemann, B.; Erre, G.; Junge, K.; Beller,<br />

M. Chem. Asian J. 2006, 1, in press.


6. Transfer hydrogenation<br />

6.1. Homogeneous iron catalyst in reduction chemistry<br />

6.1.3. Biomimetic Transfer Hydrogenation of 2–Alkoxy– and<br />

2–Aryloxyketones with Iron Porphyrin Catalysts<br />

Stephan Enthaler, Björn Spilker, Giulia Erre, Man Kin Tse, Kathrin Junge, and Matthias<br />

Beller*, Tetrahedron 2007, submitted.<br />

Contributions<br />

Ligands 1a and 1d were synthesized by Dr. M. K. Tse. Electrochemical experiments were<br />

carried out by Dr. B. Spilker.<br />

114


Pergamon<br />

Tetrahedron<br />

TETRAHEDRON<br />

Biomimetic Transfer Hydrogenation of 2-Alkoxy- and 2-<br />

Aryloxyketones with Iron Porphyrin Catalysts<br />

Stephan Enthaler, Björn Spilker, Giulia Erre, Kathrin Junge, Man Kin Tse, and Matthias Beller *<br />

Leibniz-Institut für Katalyse e.V. an der <strong>Universität</strong> <strong>Rostock</strong>, Albert-Einstein-Str. 29a, D-18059 <strong>Rostock</strong>, Germany<br />

Abstract— In situ generated iron porphyrins are applied as homogeneous catalysts in the transfer hydrogenation of αsubstituted<br />

ketones. Using 2-propanol as hydrogen donor various hydroxyl-protected 1,2-hydroxyketones are reduced to the<br />

corresponding mono-substituted 1,2-diols in good to excellent yield. Under optimized reaction conditions catalyst turnover<br />

frequencies up to 2500 h -1 have been achieved. © 2008 Elsevier Science. All rights reserved<br />

1. Introduction<br />

The relevance of 1,2-diols is impressively shown by the<br />

production of ethylene and propylene glycol on multimillion-ton<br />

scale for the synthesis of polymers and antifreeze<br />

compounds. In addition, more complex 1,2-diols and<br />

their derivatives constitute interesting building blocks for a<br />

variety of biologically active compounds.<br />

In general, synthetic approaches to 1,2-diols are based on<br />

oxidative processes of olefins. Typical processes include<br />

the epoxidation of C=C double bonds with peracids or<br />

hydroperoxides and subsequent hydrolysis, and secondly<br />

the dihydroxylation of olefins in the presence of osmium<br />

catalysts (Scheme 1). 1<br />

R 1<br />

R 4<br />

R 2<br />

R 3<br />

Dihydroxylation<br />

Epoxidation Ring-<br />

R1 O R2 opening<br />

R 4<br />

Scheme 1. Possible approaches to 1,2-diols.<br />

R 3<br />

HO OH<br />

R1 R2 R4 R3 Hydrogenation<br />

A much less explored reaction is the transition metalcatalyzed<br />

reduction of 1,2-hydroxyketones. Among the<br />

different reductions, the transfer hydrogenation with 2propanol<br />

as non-toxic hydrogen donor is practical and easy<br />

to perform. 2,3 Until now only a few number of applications<br />

were reported using Rh or Ru based metal catalysts in the<br />

reduction of hydroxy-protected 1,2-hydroxyketones, which<br />

———<br />

* Tel: +49-381-1281-113; Fax: +49-381-1281-5000; e-mail: matthias.beller@catalysis.de.<br />

R 1<br />

O<br />

R 2<br />

O<br />

R 1<br />

O<br />

R 3 R 2<br />

Reduction<br />

OR 5<br />

is in contrast to the well-studied reduction of<br />

acetophenone. 4<br />

A major challenge for catalytic hydrogenations is the<br />

substitution of expensive and rare transition metals (Ir, Rh,<br />

Ru) by ubiquitous available, inexpensive and less toxic<br />

metals. Consequently, the use of iron catalysts is especially<br />

desirable. 5 Up to now, homogeneous iron catalysts have<br />

been mostly used for carbon-carbon bond formation, such<br />

as olefin polymerizations, cross couplings, and<br />

cycloadditions as well as reduction of nitro compounds. 6<br />

However, much less attention was directed towards ironcatalyzed<br />

transfer hydrogenation. Only the groups of<br />

Noyori, 7 Vancheesan, 8 Bianchini, 9 and Gao 10 reported the<br />

utilization of iron salts and iron complexes, containing<br />

tetradentate phosphines 9 or aminophosphines 10 , in the<br />

reduction of ketones to obtain alcohols and α,β-unsaturated<br />

carbonyl compounds to yield α,β-saturated carbonyl<br />

compounds.<br />

More recently, we established two iron-based methods for<br />

the effective transfer hydrogenation of aryl alkyl as well as<br />

alkyl alkyl ketones utilizing 2-propanol as the hydrogen<br />

source. 11 On the one hand we applied a combination of iron<br />

salts, tridentate amines and phosphines as catalysts. 11a On<br />

the other hand an in situ catalyst composed of an iron salt<br />

and porphyrins has been used (Scheme 2). 11b Unfortunately,<br />

both catalyst systems faced some difficulties with αsubstituted<br />

alkyl aryl ketones. Herein we report for the first<br />

time the successful hydrogenation of α-alkoxy- and αaryloxy-substituted<br />

acetophenones in the presence of ironporphyrin-complexes.<br />

1


2<br />

R<br />

NH N<br />

N<br />

MeO 2C(H 2C) 2<br />

MeO 2C(H 2C) 2<br />

R<br />

R<br />

HN<br />

NH N<br />

N<br />

HN<br />

R<br />

(CH 2) 2CO 2Me<br />

(CH 2) 2CO 2Me<br />

1a R = p-Cl-C 6H 4<br />

1b R = C 6H 5<br />

1c R = p-C 6H 4NMe 3Tos<br />

1d R = C 6F 5<br />

1e R = 4-pyridyl<br />

HO 2C(H 2C) 2<br />

N Cl N<br />

Fe<br />

N<br />

N<br />

Tetrahedron<br />

(CH 2) 2CO 2H<br />

coproporphyrin I (2) chlorprotoporphyrin IX Fe (III) (3)<br />

Scheme 2. Selection of applied porphyrin ligands. 11b<br />

2. Results and Discussion<br />

For exploratory experiments the hydrogenation of 2methoxyacetophenone<br />

was used as a model reaction.<br />

Typically, the pre-catalyst is prepared by stirring a solution<br />

of 1 mol% iron salt and 1 mol% 1a in 2-propanol (1.0 mL)<br />

for 16 h at 65 °C. After addition of 50 mol% base the<br />

mixture is heated for 5 minutes at 100 °C and the substrate<br />

2-methoxyacetophenone 4 is added. At first, the influence<br />

of different iron precursors on the reaction rate was studied<br />

(Table 1). To our delight <strong>full</strong> conversion is obtained in all<br />

cases within 4 h. Obviously, the formation of the active<br />

catalyst is independent from the Fe-salt and is complete for<br />

all stated pre-catalysts under the standard conditions.<br />

Without any Fe present no conversion took place, also<br />

using FeCl2 without ligand gave only 37% yield. From a<br />

practical point of view we chose FeCl2 as iron source for<br />

further investigations.<br />

It is well known, that the kind and also the amount of base<br />

are important parameters for obtaining reasonable<br />

quantities of product in transfer hydrogenations. Thus, we<br />

investigated frequently used bases, such as NaOH, KOH<br />

and KO t Bu. In all cases excellent yields are achieved,<br />

except for LiOH and Cs2CO3 (Table 2, entries 1–9).<br />

Diminishing the base concentration to 5–10 mol% resulted<br />

in a decrease of yield of 2-methoxy-1-phenylethanol 5,<br />

while 25 mol% base still gave excellent yield within two<br />

hours. Noteworthy, in the absence of base no transfer of<br />

hydrogen occurred (Table 2, entry 15). In addition,<br />

nitrogen-containing bases (e.g. pyridine, NEt3, DBU) were<br />

tested, but only unsatisfactory results were attained<br />

(conversion: 99<br />

4 b)<br />

5<br />

FeCl 2<br />

FeCl 3<br />

37<br />

>99<br />

6 FeBr 2 >99<br />

7 FeI 2 >99<br />

8 Fe(acac) 2 >99<br />

9 Fe(acac) 3 >99<br />

10 FeSO 4 7H 2O >99<br />

11 CpFe(CO) 2I >99<br />

12 c)<br />

Fe(III)citrate H 2O >99<br />

13 Fe 3(CO) 12 >99<br />

Reaction conditions: 0.0038 mmol in situ catalyst (0.0038 mmol Fe-source<br />

or 0.0013 mmol Fe3(CO)12 and 0.0038 mmol 1a in 1.0 mL 2-propanol for<br />

16 h at 65 °C), 0.19 mmol NaOH in 0.5 mL 2-propanol, 0.38 mmol 2methoxyacetophenone<br />

4 in 0.5 mL 2-propanol, 4 h at 100 °C. a) Yield was<br />

determined by GC analysis (30 m HP Agilent Technologies 50–300 °C)<br />

with diglyme as internal standard. b) No ligands are added. c) Fe(III)citrate<br />

was recrystallized before use. 12<br />

On the one hand, substitution of strong coordinating<br />

chloride ligands by weaker BF4 - ligands via addition of<br />

AgBF4 was done. On the other hand activation of the<br />

catalyst by light irradiation (Perkin–Elmer PE300BF lamp<br />

with hot mirror, ~380–770 nm) has been tried.<br />

Surprisingly, both procedures led to complete deactivation<br />

of the catalysts with respect to hydrogenation reaction.<br />

In order to further improve the catalyst system we applied<br />

different porphyrin ligands (1a–3) under the optimized<br />

conditions, namely 0.5 mol% Fe catalyst or 0.01 mol%,<br />

respectively, (FeCl2 and the corresponding porphyrin in a<br />

ratio of 1:1), sodium hydroxide (25 mol% or 0.5 mol%) at<br />

100 °C (Table 3). In case of 0.5 mol% catalyst loading for<br />

all iron-porphyrin catalysts excellent conversion was<br />

attained. For 0.01 mol% catalyst loading best performance<br />

was achieved with the meso-substituted porphyrin 1a<br />

(Table 3, entry 2). Substitution at the meso-phenylic system<br />

of porphyrin with electron withdrawing groups (1c and 1d)<br />

lowered the catalyst activity. Furthermore, an excellent<br />

turnover frequency of 1850 h -1 was observed with a


porphyrin substituted at the β−pyrrolenic positions (Table<br />

3, entry 12). The complex chloroprotoporphyrin IX Fe(III)<br />

3 was utilized without catalysts pre-formation as described<br />

for all other representatives, and good turnover numbers<br />

were determined (Table 3, entries 13–14). Reducing the<br />

catalyst loading to 0.01 mol% turnover frequencies up to<br />

2100 h -1 were realized (Table 3, entries 1–3).<br />

Table 2: Influence of temperature and base in the Fe-catalyzed transfer<br />

hydrogenation of 2-methoxyacetophenone 4.<br />

O<br />

OMe<br />

FeCl 2/1a<br />

base, 2-PrOH, 4 h<br />

4 5<br />

Entry Base Base<br />

[mol%]<br />

Temp.<br />

[°C]<br />

OH<br />

OMe<br />

Yield [%] a)<br />

1 NaOH 50 100 >99<br />

2<br />

KOH 50 100 >99<br />

3 LiOH 50 100 49<br />

4 NaO i Pr 50 100 >99<br />

5 KO t Bu 50 100 >99<br />

6 b)<br />

NaO t Bu 50 100 >99<br />

7 K 2CO 3 50 100 >99<br />

8 Na 2CO 3 50 100 97<br />

9 Cs 2CO 3 50 100 37<br />

10 c)<br />

NaOH 50 25 99<br />

15 e)<br />

16 e)<br />

17 e)<br />

18 e)<br />

19 e)<br />

- - 100 99<br />

NaOH 50 100 >99<br />

Reaction conditions: 0.0038 mmol in situ catalyst (0.0038 mmol FeCl2 and<br />

0.0038 mmol 1a in 1.0 mL 2-propanol for 16 h at 65 °C), 0.19 mmol base<br />

in 0.5 mL 2-propanol, 0.38 mmol 2-methoxyacetophenone 4 in 0.5 mL 2propanol,<br />

4 h at described temperature. a) Yield was determined by GC<br />

analysis (30 m HP Agilent Technologies 50–300 °C) with diglyme as<br />

internal standard. b) A mixture of solvents was used t-BuOH and 2propanol<br />

1:4. c) Light assisted reaction, thereby only the visible range was<br />

applied since the ultraviolet range was shielded. d) Addition of 2 equiv.<br />

AgBF4 with respect to FeCl2. e) Reaction time: 2 h.<br />

Notably, the addition of small amounts of water (5 mol% to<br />

10 mol%) accelerated the transfer hydrogenation compared<br />

to the water-free conditions and turnover frequencies up to<br />

2500 h -1 were achieved. Higher amounts of water (up to<br />

500 mol%) hampered the reaction.<br />

Tetrahedron<br />

yield [%]<br />

O<br />

OMe<br />

FeCl 2/1a<br />

OH<br />

NaOH, 2-PrOH, 2 h<br />

4 5<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

2100<br />

2500<br />

2450<br />

1850<br />

1200<br />

0 5 10 25 50 500<br />

amount of water [mol%]<br />

3000<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

OMe<br />

TOF [1/h]<br />

Figure 1. The influence of water addition on hydrogenation of 2methoxyacetophenone<br />

4. Reaction conditions: 0.038 μmol in situ catalyst<br />

(0.038 μmol FeCl2 and 0.08 μmol 1a in 1.0 mL 2-propanol for 16 h at 65<br />

°C), 1.9 μmol NaOH in 0.5 mL 2-propanol, the corresponding amount of<br />

water, 0.38 mmol substrate in 0.5 mL 2-propanol, temperature: 100 °C ,<br />

reaction time: 2 h. a) Yield was determined by GC analysis (30 m HP<br />

Agilent Technologies 50–300 °C) with diglyme as internal standard.<br />

Under optimized conditions (vide supra), FeCl2, 1a, NaOH,<br />

100 °C, we compared the activity of the catalytic system<br />

for the reduction of acetophenone 6, orthomethoxyacetophenone<br />

8 and 2-methoxyacetophenone 4.<br />

Here, the yield of 1-phenylethanol 7, ortho-methoxy-1phenylethanol<br />

9 and 2-methoxy-1-phenylethanol 5 was<br />

monitored during the reaction.<br />

The attained curves are shown in Figure 2. The reduction of<br />

2-methoxyacetophenone proceeded without any induction<br />

period or deactivation process. Similar behaviour was<br />

observed for ortho-methoxyacetophenone, albeit a slight<br />

deceleration occurred after nearly 100 min. Nevertheless,<br />

<strong>full</strong> conversion was reached after 7 hours. In case of<br />

acetophenone the reduction process was significantly<br />

slower. After 8 hours a deactivation took place. Notably,<br />

the activities are in agreement with the oxidation potentials<br />

of the applied ketones (2-methoxyacetophenone, E0 = 213<br />

mV; ortho-methoxyacetophenone, E0 = 141 mV 13 ;<br />

acetophenone, E0 = 118 mV).<br />

In order to demonstrate the usefulness of our concept the<br />

catalyst containing ligand 1a was tested in the<br />

hydrogenation of several hydroxy-protected 1,2hydroxyketones<br />

(Table 4).<br />

3


4<br />

Tetrahedron<br />

Table 3: Screening of different porphyrins in the Fe-catalyzed transfer<br />

hydrogenation of 2-methoxyacetophenone 4.<br />

O<br />

OMe<br />

FeCl 2/Porphyrin<br />

OH<br />

NaOH, 2-PrOH, 2 h<br />

4 5<br />

Entry Ligand Catalyst<br />

loading [mol%]<br />

OMe<br />

Yield [%] a) TOF [h -1 ] b)<br />

1 1a 0.5 >99 >99<br />

2 1a 0.01 42 2100<br />

3 1b 0.5 >99 >99<br />

4 1b 0.01 30 1500<br />

5 1c 0.5 >99 >99<br />

6 1c 0.01 18 900<br />

7 1d 0.5 >99 >99<br />

8 1d 0.01 23 1150<br />

9 1e 0.5 >99 >99<br />

10 1e 0.01 24 1200<br />

11 2 0.5 >99 >99<br />

12 2 0.01 37 1850<br />

13 3 0.5 >99 >99<br />

14 3 0.01 26 1300<br />

Reaction conditions: 0.0019 mmol or 0.038 mmol in situ catalyst (0.0019<br />

mmol or 0.038 μmol FeCl2 and 0.0019 mmol or 0.038 μmol porphyrin in<br />

1.0 mL 2-propanol for 16 h at 65 °C), 0.095 mmol or 0.0019 mmol<br />

sodium hydroxide in 0.5 mL 2-propanol, 0.38 mmol 2methoxyacetophenone<br />

4 in 0.5 mL 2-propanol, 2 h at 100 °C. a) Yield was<br />

determined by GC analysis (30 m HP Agilent Technologies 50–300 °C)<br />

with diglyme as internal standard. b) Turnover frequencies were<br />

determined after 2 h.<br />

In general, the substrates were synthesized by reacting 2bromo<br />

ketones with different alcohols in the presence of<br />

base. 14 A first set of experiments was dedicated to the<br />

variation of the protecting group. Good to excellent yield<br />

was obtained by changing the methoxy-group by phenoxy<br />

substituents. Best results were achieved with the pchlorophenoxy<br />

substituent (Table 4, entry 2). Increasing the<br />

bulkiness at 2- and 6-position the conversion decreased<br />

significantly (Table 4, entry 4), while substitution only at 2position<br />

led to excellent conversion of the desired product.<br />

Substitution on the acetophenone skeleton by a p-methyl or<br />

p-methoxy group gave good to <strong>full</strong> conversion. In addition,<br />

indanone and pinacolin based substrates were synthesized<br />

and tested in this transfer hydrogenation, thereby moderate<br />

conversions were obtained (Table 4, entries 7 and 8).<br />

Unfortunately, variation of the hydroxyl protecting group to<br />

common acetyl and tert-butyldimethylsilyl groups<br />

displayed an inactivity of presented catalyst (Table 4,<br />

entries 11 and 12). The hydrogenation of 2-hydroxyacetophenone<br />

20 and 2-hydroxy-2-methyl-1-phenylpropan-<br />

1-one 21 were not successful since decomposition was<br />

observed. In addition, 2-morpholino-acetophenone 22,<br />

which is a promising precursor for 1,2-amino alcohols, was<br />

subjected to the transfer hydrogenation protocol.<br />

Unfortunately, no conversion was observed even if the<br />

reaction time was extended to 24 hours (Table 4, entry 15).<br />

R 1<br />

O<br />

R 2<br />

4 R1 = H; R2 = OMe<br />

100<br />

6 R1 = H; R2 = H<br />

8 R1 = 90OMe;<br />

R2 = H<br />

yield [%]<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

[Fe]/1a<br />

Na-2-OPr, 2-PrOH, 100 °C<br />

R 1<br />

OH<br />

R 2<br />

5 R 1 = H; R 2 = OMe<br />

7 R 1 = H; R 2 = H<br />

9 R 1 = OMe; R 2 = H<br />

0<br />

0 100 200 300 400 500 600 700<br />

time [min]<br />

Figure 2. Comparative study between acetophenone 6, orthomethoxyacetophenone<br />

8 and 2-methoxyacetophenone 4. Reaction<br />

conditions: 0.0038 mmol in situ catalyst (0.0038 mmol FeCl2 and 0.0038<br />

mmol 1a in 1.0 mL 2-propanol for 16 h at 65 °C), 0.19 mmol base in 0.5<br />

mL 2-propanol, 0.76 mmol substrate in 0.5 mL 2-propanol, temperature:<br />

100 °C. a) Yield was determined by GC analysis (7: 50 m Lipodex E, 95–<br />

150 °C, 7: 50 m Lipodex E, 90–180 °C, 9: 30 m HP Agilent Technologies<br />

50–300 °C) with diglyme as internal standard.<br />

In order to get further informations about the presented<br />

catalytic system we carried out electrochemical<br />

investigations of the in situ catalysts containing ligands 1ae.<br />

Cyclic voltammetry has been established as a useful tool<br />

for studying the electrochemical properties of porphyrins<br />

and metalloporphyrins. 15 The measurement conditions were<br />

as close to the reaction conditions as possible. However,<br />

under strong basic conditions, no reliable measurements<br />

were possible, due to low catalyst solubility at room<br />

temperature. Therefore measurements were performed in<br />

the absence of base. The temperature dependence of the<br />

potentials according to the Nernst equation have to be<br />

similar for all Fe-porphyrins and should be in the range of<br />

20 to 30 mV between room temperature and the reaction<br />

temperature of 100 °C. The electrochemical analysis is<br />

presented in Figure 3.<br />

All cyclic voltammograms illustrate clearly the oxidation of<br />

Fe(II) to Fe(III) in the potential range above ca. 0.2 V vs.<br />

SCE. The small (negative) reduction currents result<br />

probably from Fe(III) contamination of the FeCl2 or from<br />

oxidation of FeCl2 from other impurities (e.g. O2). From the<br />

5<br />

7<br />

9


colour of the solutions (light brown to red-brown) and the<br />

height of oxidation current (compared to 5 mmol/L FeCl2)<br />

the curves result from free FeCl2 (light yellow coloured<br />

solution) or incomplete Fe(II)-porphyrin formation can be<br />

excluded. 16<br />

Table 4: Reduction of α-substituted ketones in the presence of<br />

iron catalyst containing porphyrin 1a.<br />

O<br />

OH<br />

FeCl<br />

X<br />

2/1a<br />

R<br />

R<br />

X<br />

4 R<br />

1<br />

R1<br />

4<br />

R<br />

NaOH, 2-PrOH, 100 °C, 2 h<br />

2 R R 3<br />

2 R3 Run Substrate Product conv.<br />

[%]<br />

1<br />

2<br />

3<br />

4<br />

6<br />

7<br />

8<br />

9<br />

10<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

OMe<br />

4<br />

O<br />

10<br />

11<br />

12<br />

O<br />

i-Pr<br />

15<br />

13<br />

O<br />

14<br />

O<br />

O<br />

16<br />

t-Bu<br />

i-Pr<br />

MeO 17<br />

Cl<br />

O Ph<br />

OH<br />

OH<br />

OH<br />

OH<br />

OH<br />

OH<br />

O<br />

O<br />

OMe<br />

5<br />

O<br />

10a<br />

11a<br />

12a<br />

O<br />

i-Pr<br />

OH<br />

13a<br />

O<br />

14a<br />

O<br />

15a<br />

OH<br />

O<br />

OH<br />

16a<br />

t-Bu<br />

i-Pr<br />

MeO 17<br />

O Ph<br />

Cl<br />

Tetrahedron<br />

>99<br />

92<br />

>99 a)<br />

97<br />

74 a)<br />

43<br />

48<br />

>99 a)<br />

83 a)<br />

11<br />

12<br />

13<br />

14<br />

15<br />

Ph<br />

O<br />

O<br />

O<br />

O<br />

O<br />

OAc<br />

18<br />

O<br />

Si<br />

19<br />

OH<br />

20<br />

OH<br />

21<br />

N<br />

22<br />

O<br />

Ph<br />

OH<br />

OH<br />

OH<br />

OH<br />

OH<br />

OAc<br />

O<br />

Si<br />

19a<br />

18a<br />

OH<br />

20a<br />

OH<br />

21a<br />

N<br />

O<br />

22a<br />

5<br />


6<br />

Tetrahedron<br />

by formation of other complexes, since 2-propanol and<br />

chloride ions are present in the solution, which are possible<br />

ligands for the axial position. During oxidation of Fe(II)- to<br />

Fe(III)-porphyrin one anion has to fulfil the unsaturated<br />

situation at the Fe(III) center analogue to compound 3 in<br />

Scheme 2. At the same time it has to be considered that<br />

Fe(II)-porphyrins can also be coordinated with two axial<br />

ligands (six-coordinate Fe(II)). 15,17 The small amount of<br />

dissolved chlorid results maybe in slower kinetics of the<br />

Fe(III)-porphyrin formation, seen by the reversible<br />

potential values (Table 5).<br />

Table 5: Half-wave potentials and values for reversibility.<br />

Compound TOF E1/2 vs. SCE / mV slope / mV<br />

1d 1150 326 146<br />

1e 1200 339 141<br />

1b 1500 338 152<br />

1a 2100 330 204<br />

FeCl2 - 380 205<br />

Ferrocene - 525 61<br />

Due to the influence of the axial ligand a change or<br />

different coordination of the porphyrins at the axial<br />

positions results in a shift of potential (with an excess of<br />

chlorid the potentials shift to lower values with higher<br />

reversibility). 15 The half-wave potentials for the first<br />

oxidation step resulting from the measurements under these<br />

conditions are presented in Table 5. Under conditions<br />

identical with those for the porphyrins the<br />

ferrocenium/ferrocene couple has a half-wave potential of<br />

525 mV and a slope of 61 mV (E1/2 = 480 mV in 0.1 mol/L<br />

TBA PF6 / CH2Cl2). 18 With the half-wave potentials of the<br />

Fe-porphyrins no correlation with the TOF can be found. In<br />

addition following our standard protocol the formation of<br />

the pre-catalyst was studied by in situ mass spectroscopy.<br />

After stirring FeCl2 and ligand 1a for 16 hours at 65 °C the<br />

corresponding iron complex 19 formed by elimination of 2<br />

equiv. of HCl was detected, which is in agreement with<br />

previous reports. 20 Unfortunately, no other catalyst<br />

intermediate could be unambiguously identified. Thus, the<br />

detailed mechanism of the Fe-based hydrogenation<br />

catalysts is still unclear.<br />

In conclusion, we have demonstrated the successful<br />

extension of the Fe-catalyzed transferhydrogenation to α-<br />

hydroxy-protected hydroxyketones. Under optimized<br />

conditions turnover frequencies up to 2500 h -1 were<br />

achieved. The scope and limitation of the catalyst were<br />

demonstrated on reduction of ten different ketones with<br />

good to excellent yields.<br />

3. Experimental Section<br />

3.1. General<br />

1 H and 13 C NMR spectra were recorded on Bruker<br />

Spectrometer 400 and 300 ( 1 H: 400.13 MHz, and 300.13<br />

MHz; 13 C: 100.6 MHz, and 75.5 MHz). The calibration of<br />

1 H and 13 C spectra was carried out either on solvent signals<br />

(δ CDCl3) = 7.25 and 77.0) or TMS. Mass spectra were<br />

recorded on an AMD 402 spectrometer. IR spectra were<br />

recorded as KBr pellets or Nujol mulls on a Nicolet Magna<br />

550. All manipulations were performed under argon<br />

atmosphere using standard Schlenk techniques. Unless<br />

specified, all chemicals are commercially available and<br />

used as received. 2-Propanol was used without further<br />

purification (purchased from Fluka, dried over molecular<br />

sieves). Sodium 2-propylate and sodium t-butylate were<br />

prepared by reacting sodium with 2-propanol or t-butanol,<br />

respectively, under an argon atmosphere (stock solution).<br />

Ketone 8 was dried over CaH2, distilled in vacuum and<br />

stored under argon. Substrates 18 and 19 were used without<br />

further purifications. Porphyrins 1a and 1d were<br />

synthesized according to literature protocols. 21 Fe(III)citrate<br />

was recrystallized from water/ethanol before use since it<br />

was purchased in technical grade.<br />

3.2. General procedure for the synthesis of substrates:<br />

A solution of 2-bromo ketone (10.1 mmol), phenol (10.1<br />

mmol), K2CO3 (15 mmol) in acetone (15 mL) was refluxed<br />

for 3 hours. The solvent was removed in vacuum and the<br />

residue dissolved in ethyl acetate / diethyl ether and washed<br />

with water and brine. After drying over Na2SO4 the solvent<br />

was removed and the crude product was purified by<br />

crystallization.<br />

2-Phenoxy-acetophenone (10): 10.1 mmol scale, flash<br />

column chromatography (eluent: ethyl acetate/n-hexane<br />

4:1), recrystallized from 2-propanol. Yield: 39% (white<br />

crystals). Mp = 60–61 °C. 1 H NMR (300 MHz, CDCl3) δ =<br />

8.02–7.98 (m, 2H); 7.64–7.58 (m, 1H); 7.53–7.46 (m, 2H);<br />

7.31–7.22 (m, 2H); 7.01–6.90 (m, 3H); 5.28 (s, 2H, CH2).<br />

13 C NMR (75.5 MHz, CDCl3): δ = 194.5; 157.9; 134.5;<br />

133.8; 129.5; 128.8; 128.1; 121.6; 114.7; 70.7. IR (KBr):<br />

3387 w; 3060 w; 2897 w; 2843 w; 2749 w; 1928 w; 1708 s;<br />

1599 s; 1499 s; 1480 m; 1449 m; 1433 m; 1386 w; 1341 w;<br />

1303 m; 1292 m; 1250 s; 1228 s; 1189 m; 1175 m; 1094 m;<br />

1076 m; 1028 w; 1001 m; 975 m; 921 w; 886 w; 872 m;<br />

751 s; 688 s; 665 m; 614 w; 584 w; 555 w; 511 m; 414 w.<br />

MS (EI): m/z (%) = 212 ([M + ], 26); 105 (100); 77 (43); 51<br />

(12). HRMS calculated for C14H14O2: 214.09883; Found:<br />

214.098506.<br />

2-(4-Chlorophenoxy)-acetophenone (11): 7.8 mmol scale,<br />

refluxing time: 12 h, crystallized from 2-propanol. Yield:<br />

73% (white plates). Mp = 87–88 °C. 1 H NMR (300 MHz,<br />

CDCl3) δ = 8.01–7.96 (m, 2H); 7.66–7.56 (m, 1H); 7.54–<br />

7.47 (m, 2H); 7.26–7.20 (m, 2H); 6.90–6.83 (m, 2H); 5.26


(s, 2H, CH2). 13 C NMR (75.5 MHz, CDCl3): δ = 194.0;<br />

156.6; 134.3; 134.0; 129.4; 128.9; 128.0; 126.5; 116.1;<br />

70.9. IR (KBr): 3375 w; 3060 w; 2901 w; 2851 w; 2737 w;<br />

1699 s; 1596 s; 1581 m; 1490 s; 1448 m; 1437 s; 1409 w;<br />

1385 w; 1338 w; 1317 w; 1289 s; 1231 s; 1172 m; 1106 m;<br />

1096 m; 1089 m; 1075 m; 1001 m; 983 s; 921 w; 870 w;<br />

830 s; 813 m; 786 w; 756 s; 701 w; 686 s; 634 w; 587 m;<br />

508 m; 476 w; 457 w; 441 w; 422 w; 406 w. MS (EI): m/z<br />

(%) = 246 ([M + ], 15); 105 (100); 77 (30). HRMS calculated<br />

for C14H11ClO2: 246.04421; Found: 246.044678.<br />

2-(2-tert-Butyl-4-methylphenoxy)-acetophenone (12): 6.0<br />

mmol scale, refluxing time: 12 h, crystallized from 2propanol.<br />

Yield: 44% (white needles). Mp = 94–95 °C. 1 H<br />

NMR (300 MHz, CDCl3) δ = 8.05–8.00 (m, 2H); 7.66–7.59<br />

(m, 1H); 7.55–7.48 (m, 2H); 7.13 (d, 1H, J = 2.26 Hz); 6.96<br />

(ddd, 1H, J = 8.10 Hz, J = 2.26 Hz, J = 0.75 Hz); 6.76 (d,<br />

1H, J = 8.22 Hz); 5.29 (s, 2H, CH2); 2.30 (s, 3H, CH3);<br />

1.43 (s, 9H, C(CH3)3). 13 C NMR (75.5 MHz, CDCl3): δ =<br />

194.5; 154.8; 138.4; 134.7; 133.7; 130.3; 128.8; 128.1;<br />

127.8; 127.1; 112.3; 71.0; 34.7; 29.9; 20.8. IR (KBr): 3399<br />

w; 3051 w; 2958 s; 2901 m; 2856 m; 2733 w; 1709 s; 1598<br />

m; 1581 w; 1499 s; 1448 m; 1439 m; 1404 w; 1386 m;<br />

1357 w; 1287 m; 1266 m; 1244 m; 1226 s; 1181 w; 1156<br />

w; 1107 m; 1074 w; 1020 w; 1001 m; 980 s; 930 w; 919 w;<br />

876 w; 861 m; 802 s; 751 s; 687 s; 655 m; 584 m; 491 m.<br />

MS (EI): m/z (%) = 282 ([M + ], 64); 267 (21); 249 (21); 161<br />

(11); 121 (13); 119 (16); 117 (14); 105 (100); 91 (33); 77<br />

(29). HRMS calculated for C19H22O2: 28216143; Found:<br />

282.162186.<br />

2-(2,6-Di-iso-propylphenoxy)-acetophenone (13): 5.6<br />

mmol scale, refluxing time: 12 h, crystallized from 2propanol.<br />

Yield: 34% (yellow crystals). Mp = 60–65 °C. 1 H<br />

NMR (300 MHz, CDCl3) δ = 8.01–7.95 (m, 2H); 7.66–7.59<br />

(m, 1H); 7.55–7.47 (m, 2H); 7.16 (s, 3H); 5.12 (s, 2H,<br />

CH2); 3.34 (sept, 2H, J = 6.91 Hz, CH); 1.25 (d, 12H, J =<br />

6.87 Hz, CH3). 13 C NMR (75.5 MHz, CDCl3): δ = 193.9;<br />

153.0; 141.6; 133.7; 128.8; 127.8; 125.2; 124.3; 76.7; 26.5;<br />

24.1. IR (KBr): 3442 br; 3063 w; 3028 w; 2967 s; 2868 m;<br />

1706 s; 1597 m; 1580 w; 1451 m; 1429 m; 1381 w; 1371<br />

w; 1360 m; 1334 m; 1255 m; 1227 m; 1185 s; 1163 m;<br />

1100 m; 1085 m; 1075 w; 1058 w; 1042 w; 1001 w; 992 w;<br />

972 m; 934 w; 854 w; 801 m; 763 m; 755 s; 692 m; 648 w;<br />

617 w; 586 w; 556 w; 521 w; 472 w; 434 w. MS (EI): m/z<br />

(%) = 296 ([M + ], 10); 176 (83); 161 (100); 147 (20); 133<br />

(16); 105 (59); 91 (44); 77 (28); 43 (11). HRMS calculated<br />

for C20H24O2: 296.17708; Found: 296.176483.<br />

2-Phenoxy-2,3-dihydro-1H-inden-1-one (14): 9.5 mmol<br />

scale, refluxing time: 16 h, Yield: 56% (off-white crystals).<br />

Mp = 75–78 °C. 1 H NMR (300 MHz, CDCl3) δ = 7.85 (d,<br />

1H, J = 7.72 Hz); 7.67 (dt, 1H, J = 7.44 Hz, J = 1.13 Hz);<br />

7.50–7.41 (m, 2H); 7.37–6.99 (m, 2H); 7.10–6.99 (m, 3H);<br />

5.09 (dd, 1H, J = 7.54 Hz, J = 4.52 Hz, CH); 3.73 (dd, 1H,<br />

J = 7.54 Hz, J = 16.95 Hz, CH2); 3.19 (dd, 1H, J = 16.95<br />

Hz, J = 4.33 Hz, CH2). 13 C NMR (75.5 MHz, CDCl3): δ =<br />

201.7; 157.9; 150.6; 135.9; 134.6; 129.5; 128.2; 126.7;<br />

Tetrahedron<br />

124.6; 121.7; 115.7; 77.8; 34.1. IR (KBr): 3413 br; 3063 w;<br />

3039 w; 3028 w; 2916 w; 2906 w; 1716 s; 1608 m; 1597<br />

m; 1587 m; 1493 m; 1474 m; 1464 m; 1431 w; 1325 w;<br />

1299 m; 1276 m; 1250 m; 1233 s; 1208 w; 1174 w; 1156<br />

w; 1080 m; 1070 m; 1038 w; 1020 w; 1005 m; 956 w; 913<br />

m; 889 w; 753 s; 725 m; 691 m; 618 w; 607 w; 502 w; 455<br />

w. MS (EI): m/z (%) = 224 ([M + ], 53); 131 (100); 103 (32);<br />

94 (11); 77 (25). HRMS calculated for C15H12O2:<br />

224.08318; Found: 224.083160.<br />

3,3-Dimethyl-1-phenoxybutan-2-one (15): 11.2 mmol<br />

scale, refluxing time: 16 h. Yield: 92% (off-white crystals).<br />

Mp = 33–36 °C. 1 H NMR (300 MHz, CDCl3) δ = 7.32–<br />

7.24 (m, 2H); 7.00–6.93 (m, 1H); 6.91–6.84 (m, 2H); 4.87<br />

(s, 2H, CH2); 1.25 (s, 9H, C(CH3)3). 13 C NMR (75.5 MHz,<br />

CDCl3): δ = 209.4; 158.0; 129.5; 121.4; 114.6; 68.8; 43.1;<br />

26.3. IR (KBr): 3420 br; 3105 w; 3065 w; 3043 w; 2973 m;<br />

2931 m; 2872 w; 2846 w; 2747 w; 1721 s; 1678 w; 1599<br />

m; 1587 m; 1494 s; 1461 m; 1446 m; 1431 m; 1386 w;<br />

1365 m; 1329 w; 1291 m; 1234 s; 1179 m; 1152 w; 1113<br />

m; 1078 w; 1052 s; 1026 w; 1000 m; 990 s; 962 m; 935 w;<br />

880 m; 825 m; 766 s; 759 s; 693 m; 587 w; 545 w; 522 m.<br />

MS (EI): m/z (%) = 192 ([M + ], 33); 108 (13); 77 (26); 57<br />

(100); 41 (16). HRMS calculated for C12H16O2: 192.11448;<br />

Found: 192.114302.<br />

2-Phenoxy-4´-methylacetophenone (16): 9.4 mmol scale,<br />

refluxing time: 16 h, crystallized from 2-propanol. Yield:<br />

51% (colourless crystals). Mp = 62–63 °C. 1 H NMR (300<br />

MHz, CDCl3) δ = 7.94–7.88 (m, 2H); 7.33–7.24 (m, 4H);<br />

7.02–6.91 (m, 3H); 5.25 (s, 2H, CH2); 2.43 (s, 3H, CH3).<br />

13 C NMR (75.5 MHz, CDCl3): δ = 194.1; 158.0; 144.8;<br />

132.1; 129.54; 129.48; 128.2; 121.6; 114.6; 70.7; 21.8. IR<br />

(KBr): 3441 br, 3063 w; 3036 w; 2919 w; 2902 w; 2847 w;<br />

1695 s; 1606 s; 1587 m; 1576 m; 1498 s; 1455 w; 1433 m;<br />

1411 m; 1385 w; 1292 m; 1254 m; 1236 s; 1208 m; 1192<br />

m; 1174 m; 1149 w; 1124 w; 1113 w; 1092 m; 1075 w;<br />

1044 w; 1027 w; 1001 m; 979 m; 885 m; 871 m; 816m;<br />

749 s; 711 w; 690 m; 609 w; 584 m; 551 w; 517 w; 498 w;<br />

461 w. MS (EI): m/z (%) = 226 ([M + ], 15); 119 (100); 91<br />

(24); 77 (10). HRMS calculated for C15H14O2: 226.09883;<br />

Found: 226.099202.<br />

2-Phenoxy-4´-methoxyacetophenone (17): 8.7 mmol<br />

scale, refluxing time: 16 h, crystallized from 2-propanol.<br />

Yield: 74% (off-white needles). Mp = 54–55 °C. 1 H NMR<br />

(300 MHz, CDCl3) δ = 7.96–7.89 (m, 2H); 7.24–7.16 (m,<br />

2H); 6.93–6.84 (m, 5H); 5.13 (s, 2H, CH2); 3.80 (s, 3H,<br />

CH3). 13 C NMR (75.5 MHz, CDCl3): δ = 193.1; 164.0;<br />

158.0; 130.5; 129.0; 127.6; 121.5; 114.7; 114.0; 70.7; 55.5.<br />

IR (KBr): 3442 br; 3056 w; 3006 w; 2955 w; 2934 w; 2840<br />

w; 1960 m; 1687 s; 1653 w; 1601 s; 1511 m; 1497 s; 1459<br />

m; 1419 w; 1368 m; 1314 m; 1265 s; 1245 m; 1226 s; 1171<br />

s; 1154 w; 1116 m; 1075 m; 1032 m; 976 s; 884 m; 835 m;<br />

789 m; 769 m; 751 s; 691 m; 631 w; 608 m; 582 m; 543 w;<br />

511 w; 492 w. MS (EI): m/z (%) = 242 ([M + ], 10); 135<br />

(100); 77 (18). HRMS calculated for C15H14O3: 242.09375;<br />

Found: 242.093706.<br />

7


8<br />

Tetrahedron<br />

2-Oxo-2-phenylethyl acetate (18): 2–Hydroxy–1–<br />

phenylethanone (8.0 mmol) was dissolved in<br />

dichloromethane (10 mL) and acetic anhydride (8.0 mmol),<br />

pyridine (8.5 mmol) and 4–dimethylaminopyridine (0.08<br />

mmol) were added. The solution was stirred for 2 hours<br />

under refluxing conditions. The mixture was washed with<br />

water and brine. After drying over Na2SO4 the solvent was<br />

removed and yellow crystals were obtained. Yield: 83%.<br />

Mp = 48–50 °C. 1 H NMR (300 MHz, CDCl3) δ = 7.93–<br />

7.88 (m, 2H); 7.63–7.56 (m, 1H); 7.51–7.44 (m, 2H); 5.34<br />

(s, 2H, CH2); 2.22 (s, 3H, CH3). 13 C NMR (75.5 MHz,<br />

CDCl3): δ = 192.0; 170.3; 134.1; 133.8; 128.7; 127.6; 65.9;<br />

20.4. IR (KBr): 3460 w; 3375 w; 3064 m; 2982 w; 2937 m;<br />

1741 s; 1699 s; 1599 m; 1581 w; 1493 w; 1452 m; 1423 m;<br />

1374 m; 1321 w; 1280 m; 1244 s; 1228 s; 1185 m; 1087 m;<br />

1077 m; 1047 m; 1020 m; 996 m; 970 m; 849 m; 812 m;<br />

757 m; 688 m; 634 m; 597 m; 566 m; 483 m; 443 w; 414<br />

w. MS (EI): m/z (%) = 178 ([M + ], 1); 105 (100); 77 (34);<br />

51 (10); 43 (14). HRMS calculated for C15H14O3:<br />

178.06245; Found: 178.063004.<br />

2-(tert-Butyldimethylsilyloxy)-1-phenylethanone (19): 2–<br />

Hydroxy–1–phenylethanone (8.0 mmol) was dissolved in<br />

dichloromethane (10 mL) and imidazole (8.5 mmol) was<br />

added. After stirring for 10 minutes at room temperature<br />

tert–butyldimethylchlorosilane (8.5 mmol) was added<br />

dropwise, while rapidly a white precipitate was formed.<br />

The mixture was stirred for 5 hours at room temperature.<br />

Water was added and the organic layer was washed with<br />

water and brine. After removal of the solvent a yellow oil<br />

was obtained. Yield: 89%. 1 H NMR (300 MHz, CDCl3) δ =<br />

7.95–7.89 (m, 2H); 7.61–7.42 (m, 3H); 4.93 (s, 2H, CH2);<br />

0.94 (s, 9H, C(CH3)3); 0.13 (s, 6H, Si(CH3)2). 13 C NMR<br />

(75.5 MHz, CDCl3): δ = 197.4; 134.9; 132.2; 128.6; 127.9;<br />

67.4; 25.8; 18.5; -5.4. IR (KBr): 2953 w; 2929 m; 2885 w;<br />

2856 m;1705 s; 1599 m; 1581 w; 1472 m; 1449 m; 1390 w;<br />

1362 w; 1288 m; 1254 m; 1229 m; 1151 s; 1025 w; 1002<br />

m; 974 s; 939 w; 834 s; 777 s; 753 s; 712 m; 688 s; 670 s.<br />

MS (EI): m/z (%) = 251 ([M + +H], 16); 193 (66); 181 (10);<br />

149 (12); 135 (13); 105 (100); 75 (73).<br />

2-Morpholino-acetophenone (22): Morpholine (62 mmol)<br />

was added dropwise to a solution of 2–bromoacetophenone<br />

(7.5 mmol) in THF (15 mL). A white precipitate was<br />

formed while the mixture was refluxed for one day. An<br />

aqueous solution of NaHCO3 was added. After extraction<br />

with diethyl ether the organic layer was washed with<br />

aqueous solution of NaOH and dried over Na2SO4. The<br />

crude product was purified by flash column<br />

chromatography (eluent: ethyl acetate/n–hexane 1:1) to<br />

yield a brownish oil. Yield: 89%. 1 H NMR (300 MHz,<br />

CDCl3) δ = 8.02–7.97 (m, 2H); 7.60–7.55 (m, 1H); 7.50–<br />

7.43 (m, 2H); 3.83 (s, 2H, C(O)CH2); 3.79 (m, 4H, CH2);<br />

2.62 (m, 4H, CH2). 13 C NMR (75.5 MHz, CDCl3): δ =<br />

172.0; 160.9; 131.9; 129.6; 128.1; 67.1; 66.4; 64.4; 41.7;<br />

40.6. IR (KBr): 3415 w; 3060 m; 2970 m; 2852 m; 2762 w;<br />

1924 w; 1674 s; 1598 s; 1558 m; 1493 w; 1449 s; 1384 s;<br />

1314 m; 1300 m; 1271 s; 1230 m; 1176 m; 1113 s; 1069 m;<br />

1047 w; 1022 m; 1006 m; 981 w; 931 w; 876 w; 840 w;<br />

812 w; 788 w; 757 w; 722 m; 693 w; 674 w; 593 w; 547 w;<br />

449 w. MS (EI): m/z (%) = 205 ([M + ], 2); 100 (100); 77<br />

(18); 56 (21). HRMS calculated for C12H15O2N: 205.10973;<br />

Found: 205.109149. Rf (ethyl acetate/n-hexane 1:1) = 0.14.<br />

General procedure for the catalytic transfer<br />

hydrogenation: In a 10 mL Schlenk tube, the in situ<br />

catalyst (0.0038 mmol) is prepared stirring a solution of<br />

FeCl2 (0.0038 mmol) and porphyrin (0.0038 mmol) in 1.0<br />

mL 2-propanol for 16 h at 65 °C. The pre-catalyst system is<br />

reacted with sodium hydroxide (0.38 mmol in 0.5 mL 2propanol)<br />

and 2-methoxyacetophenone (0.38 mmol in 0.5<br />

mL 2-propanol) for 2 h at 100 °C. The solution is cooled to<br />

r.t. and filtered over a plug of silica. The conversion was<br />

measured by GC without further manipulations and the<br />

products were isolated by column chromatography or<br />

crystallization.<br />

2-Phenoxy-1-phenylethanol (10a): Mp = 48–50 °C (white<br />

crystals). 1 H NMR (300 MHz, CDCl3) δ = 7.40–7.15 (m,<br />

7H, Ph); 6.92–6.80 (m, 3H, Ph); 5.03 (dd, 1H, J = 8.67 Hz,<br />

J = 3.20 Hz, CH); 4.03–3.88 (m, 2H, CH2); 2.55 (br, 1H,<br />

OH). 13 C NMR (75.5 MHz, CDCl3): δ = 158.3; 139.6;<br />

129.5; 128.5; 128.2; 126.3; 121.3; 114.6; 73.2; 72.5. IR<br />

(KBr): 3302 br; 3060 w; 3028 w; 2932 w; 2877 w; 1598 s;<br />

1585 M; 1498 s; 1453 M; 1389 w; 1346 w; 1303 m; 1292<br />

m; 1249 s; 1196 m; 1172 m; 1152 w; 1098 m; 1078 m;<br />

1067 m; 1046 m; 1028 m; 995 w; 917 m; 884 w; 863 m;<br />

792 w; 754 s; 701 s; 692 s; 637 m; 615 m; 594 m; 541 m;<br />

512 m. MS (EI): m/z (%) = 214 ([M + ], 12); 108 (100); 94<br />

(46); 79 (60); 51 (18). HRMS calculated for C14H14O2:<br />

214.09883; Found: 214.098407.<br />

2-(4-Chlorophenoxy)-1-phenylethanol (11a): Mp = 46–<br />

48 °C (white crystals). 1 H NMR (300 MHz, CDCl3) δ =<br />

7.48–7.19 (m, 7H); 6.87–6.80 (m, 2H); 5.11 (dd, 1H, J =<br />

8.48 Hz, J = 3.38 Hz, CH); 4.08–3.95 (m, 2H, CH2); 2.70<br />

(br, 1H, OH). 13 C NMR (75.5 MHz, CDCl3): δ = 157.0;<br />

139.4; 129.4; 128.6; 128.3; 126.2; 126.1; 115.8; 72.6; 72.5.<br />

IR (KBr): 3313 br; 3059 w; 3032 w; 2933 w; 2875 w; 1596<br />

m; 1581 m; 1491 s; 1453 m; 1388 w; 1345 w; 1290 m;<br />

1242 s; 1196 w; 1169 m; 1093 m; 1066 m; 1039 m; 1006<br />

m; 914 m; 864 w; 825 m; 801 m; 749 m; 700 m; 671 m;<br />

633 w; 616 m; 559 w; 507 m. MS (EI): m/z (%) = 248<br />

([M + ], 21); 142 (34); 128 (70); 107 (100); 91 (12); 79 (31).<br />

HRMS calculated for C14H13ClO2: 248.05986; Found:<br />

248.059204.<br />

2-(2-tert-Butyl-4-methylphenoxy)-1-phenylethanol<br />

(12a): Mp = 61–64 °C (colourless crystals). 1 H NMR (300<br />

MHz, CDCl3) δ = 7.53–7.32 (m, 5H, Ph); 7.13 (d, 1H, J =<br />

2.13 Hz, Ph); 6.98 (ddd, 1H, J = 8.29 Hz, J = 2.26 Hz, J =<br />

0.75 Hz, Ph); 6.78 (d, 1H, J = 8.09 Hz, Ph); 5.21 (dd, 1H, J<br />

= 7.35 Hz, J = 4.71 Hz, CH); 4.18–4.08 (m, 2H, CH2); 2.67<br />

(br, 1H, OH); 2.31 (s, 3H, CH3); 1.41 (s, 9H, C(CH3)3). 13 C<br />

NMR (75.5 MHz, CDCl3): δ = 155.0; 139.9; 137.7; 129.9;<br />

128.5; 128.1; 127.7; 127.1; 126.3; 112.2; 73.5; 72.9; 34.7;


30.0; 20.8. IR (KBr): 3570 m; 3030 w; 2994 w; 2856 m;<br />

2917 m; 2867 m; 1605 w; 1581 w; 1497 s; 1451 s; 1406 m;<br />

1389 m; 1359 m; 1327 w; 1294 m; 1264 m; 1228 s; 1198<br />

m; 1151 m; 1095 m; 1060 m; 1033 s; 1001 w; 934 w; 911<br />

m; 881 w; 854 m; 809 m; 764 m; 702 s; 624 w; 602 w; 592<br />

w; 553 w; 516 w; 493 w. MS (EI): m/z (%) = 284 ([M + ],<br />

27); 164 (31); 149 (100); 121 (18); 107 (17); 91 (21); 77<br />

(16). HRMS calculated for C19H24O2: 284.17708; Found:<br />

284.176978.<br />

2-(2,6-Di-iso-propylphenoxy)-1-phenylethanol (13a): Mp<br />

= 67–69 °C (colourless crystals). 1 H NMR (300 MHz,<br />

CDCl3) δ = 7.46–7.26 (m, 5H); 7.09 (s, 3H); 5.16 (dd, 1H,<br />

J = 7.53 Hz, J = 4.40 Hz, CH); 3.92–3.82 (m, 2H, CH2);<br />

3.29 (sept, 2H, J = 6.91 Hz, CH(CH3)2); 3.10 (br, 1H); 1.22<br />

(dd, 12H, J = 6.90 Hz, J = 1.29 Hz). 13 C NMR (75.5 MHz,<br />

CDCl3): δ = 152.4; 141.6; 139.8; 128.4; 128.0; 126.1;<br />

125.8; 124.9; 79.3; 73.3; 26.4; 24.04; 24.00. IR (KBr):<br />

3064 m; 3030 m; 2963 s; 2927 s; 2868 m; 1604 w; 1589 w;<br />

1454 s; 1384 m; 1362 m; 1327 m; 1255 m; 1182 s; 1100 m;<br />

1047 s; 1020 s; 936 w; 913 m; 861 w; 834 w; 802 m; 755<br />

m; 700 s; 682 w; 623 w; 590 w; 527 w. MS (EI): m/z (%) =<br />

298 ([M + ], 8); 178 (51); 163 (100); 107 (15); 91 (16); 77<br />

(11). HRMS calculated for C20H26O2: 298.19273; Found:<br />

298.192353.<br />

2-Phenoxy-2,3-dihydro-1H-inden-1-ol (14a): colourless<br />

crystals. During the reaction a mixture of diastereomers<br />

were formed in a ratio of 1:3 (D1:D2). 1 H NMR (400 MHz,<br />

CDCl3) δ = 7.56–6.93 (m); 5.34 (d, 1H, J = 4.12 Hz, D1);<br />

5.28 (d, J = 5.22 Hz, D2); 5.08–5.02 (m, 1H, D2); 4.98–<br />

4.89 (m, 1H, D1); 3.53 (dd, 1H, J = 16.37 Hz, J = 6.92 Hz,<br />

D1); 3.20 (d, 2H, J = 4.40 Hz, D2); 2.98 (dd, 1H, J = 16.37<br />

Hz, J = 5.32 Hz, D1); 2.81 (br, OH). 13 C NMR (100.6<br />

MHz, CDCl3): δ = 157.7; 142.4; 139.5; 139.2; 129.6;<br />

129.5; 128.9; 128.6; 127.3; 125.1; 125.0; 124.8; 124.5;<br />

121.7; 121.6; 121.0; 115.8; 115.5; 85.4; 80.3; 75.6; 36.4;<br />

35.9. IR (Nujol): 3182 w; 2723 w; 1598 m; 1585 m; 1496<br />

m; 1459 s; 1377 s; 1291 m; 1250 s; 1207 w; 1170 w; 1155<br />

w; 1118 m; 1073 w; 1050 w; 1020 w; 1001 w; 879 w; 849<br />

w; 779 w; 746 s; 723 m; 693 m; 639 w; 511 w; 413 w. MS<br />

(EI): m/z (%) = 226 ([M + ], 19); 133 (100); 115 (21); 103<br />

(21); 94 (61); 77 (29); 65 (11). HRMS calculated for<br />

C15H14O2: 226.09883; Found: 226.099031.<br />

3,3-Dimethyl-1-phenoxybutan-2-ol (15a): Mp = 31–33 °C<br />

(colourless crystals). 1 H NMR (300 MHz, CDCl3) δ =<br />

7.33–7.24 (m, 2H); 7.00–6.88 (m, 3H); 4.14 (dd, 1H, J =<br />

9.23 Hz, J = 2.45 Hz, CH); 3.86 (m, 2H, CH2); 2.35 (br,<br />

1H, OH); 1.01 (s, 9H, C(CH3)3). 13 C NMR (75.5 MHz,<br />

CDCl3): δ = 158.6; 129.5; 121.1; 114.6; 77.2; 69.3; 33.5;<br />

26.0. IR (KBr): 3272 br; 3035 w; 2954 s; 2876 m; 1601 m;<br />

1586 m; 1499 s; 1459 m; 1393 w; 1366 m; 1338 m; 1296<br />

m; 1245 s; 1190 w; 1174 m; 1152 w; 1095 m; 1079 m;<br />

1044 m; 1020 m; 1020 m; 1005 m; 992 w; 942 w; 926 m;<br />

902 m; 895 m; 821 w; 754 s; 692 m; 598 w; 572 w; 513 m;<br />

402 w. MS (EI): m/z (%) = 194 ([M + ], 26); 119 (10); 108<br />

(63); 94 (100); 87 (19); 77 (22); 69 (14); 57 (24); 41 (17).<br />

Tetrahedron<br />

HRMS calculated for C12H18O2: 194.13013; Found: 194.<br />

129870.<br />

2-Phenoxy-1-(4´-methylphenyl)ethanol (16a): Mp = 50–<br />

51 °C (colourless crystals). 1 H NMR (300 MHz, CDCl3) δ<br />

= 7.34–7.14 (m, 6H); 6.98–6.85 (m, 3H); 5.05 (dd, 1H, J =<br />

8.64 Hz, J = 3.36 Hz, CH); 4.04 –3.97 (m, 2H, CH2); 2.87<br />

(br, 1H, OH); 2.34 (s, 3H, CH3). 13 C NMR (75.5 MHz,<br />

CDCl3): δ = 158.3; 137.8; 136.7; 129.5; 129.2; 126.2;<br />

121.1; 114.5; 73.2; 72.3; 21.1. IR (KBr): 3297 br; 3027 w;<br />

2912 w; 2880 w; 1932 w; 1599 m; 1586 m; 1514 m; 1498<br />

s; 1460 m; 1388 m; 1340 m; 1293 m; 1250 s; 1195 m; 1180<br />

m; 1172 m; 1151 w; 1090 s; 1045 m; 1021 m; 995 w; 975<br />

w; 943 w; 916 m; 886 w; 868 m; 838 w; 815 m; 756 s; 716<br />

w; 693 m; 616 w; 597 m; 575 w; 537 m; 513 m; 487 w; 464<br />

w; 403 w. MS (EI): m/z (%) = 228 ([M + ], 6); 121 (100);<br />

108 (68); 105 (15); 65 (11). HRMS calculated for<br />

C15H16O2: 228.11448; Found: 228.114429.<br />

2-Phenoxy-1-(4´-methoxyphenyl)ethanol (17a): yellow<br />

oil. 1 H NMR (300 MHz, CDCl3) δ = 7.32–7.15 (m, 4H);<br />

6.92–6.77 (m, 5H); 4.98 (dd, 1H, J = 8.67 Hz, J = 3.39 Hz,<br />

CH); 4.00–3.87 (m, 2H, CH2); 3.73 (s, 3H, CH3); 2.77 (br,<br />

1H). 13 C NMR (75.5 MHz, CDCl3): δ = 159.4; 158.3;<br />

131.7; 129.5; 127.5; 121.2; 114.6; 113.9; 73.2; 72.1; 55.3.<br />

IR (KBr): 3440 br; 3072 w; 3044 w; 2999 w; 2929 m; 2831<br />

w; 1728 w; 1612 m; 1600 m; 1587 m; 1514 s; 1497 s; 1456<br />

m; 1303 m; 1245 s; 1174 m; 1154 w; 1114 w; 1080 m;<br />

1036 m; 912 w; 867 w; 832 m; 755 m; 692 m; 639 w; 596<br />

w; 552 w; 511 w. MS (EI): m/z (%) = 244 ([M + ], 2); 137<br />

(100); 121 (10); 108 (23); 94 (12); 77 (21). HRMS<br />

calculated for C15H16O3: 244.10940; Found: 244.109218.<br />

3.3. Electrochemical measurements<br />

The electrochemical workstation is composed of an<br />

Autolab PGSTAT 10 potentiostat from ECO Chemie and a<br />

P-450 PC with GPES 4.9 software. A three-electrode<br />

arrangement with a platinum microelectrode (25 µm<br />

diameter) as the working electrode, a saturated calomel<br />

electrode (SCE) as reference electrode and a platinum<br />

counter electrode (16 mm 2 area) was used. All potentials in<br />

the following are cited against SCE. All measurements<br />

were carried out in 2-PrOH at room temperature. As<br />

supporting electrolyte we used 0.1 mol/L tetra-nbutylammonium<br />

tetrafluoraborate. Due to the low solubility<br />

at r.t. all solutions of porphyrins with FeCl2 were saturated<br />

solutions. For some porphyrins (e.g. compound 3)<br />

solubility was too low to see any currents. Reversibility<br />

was determined by plotting E vs. log[(Ia-I)/I-Ic] for the<br />

nearly linear range of the forward scan and determining the<br />

slope for the quasi-reversible electrode reactions. Anodic<br />

diffusion current (Ia) and cathodic diffusion current (Ic) are<br />

determined from the results of a sigmoidal fit with<br />

Microcal Origin 7.5.<br />

Acknowledgments<br />

9


10<br />

Tetrahedron<br />

We thank Mrs. S. Buchholz, and Dr. C. Fischer (both<br />

Leibniz–Institut für Katalyse e.V. an der <strong>Universität</strong><br />

<strong>Rostock</strong>) for excellent analytical assistance and Prof.<br />

Jeroschewski for general discussions. Generous financial<br />

support from the state of Mecklenburg–Western Pomerania<br />

and the BMBF as well as the Deutsche<br />

Forschungsgemeinschaft (Leibniz–price) are grate<strong>full</strong>y<br />

acknowledged.<br />

Note and References<br />

1. (a) Noyori, R. (Ed.) Asymmetric Catalysis in Organic Synthesis,<br />

Wiley, New York, 1994; (b) Cornils B.; Herrmann, W. A. (Eds.)<br />

Applied Homogeneous Catalysis with Organometallic Compounds,<br />

Wiley-VCH, Weinheim, 1996; (c) Jacobsen, E. N.; Pfaltz, A.;<br />

Yamamoto, H. (Eds.), Comprehensive Asymmetric Catalysis,<br />

Springer, Berlin, 1999; (d) Kitamura, M.; Noyori, R. in Murahashi,<br />

S.-I. (Ed.) Ruthenium in Organic Synthesis, Wiley-VCH, Weinheim,<br />

2004; (e) Beller, M.; Bolm, C. (Eds.), Transition Metals for Organic<br />

Synthesis, Wiley-VCH, Weinheim, 2 nd Ed. 2004; (f) J.-E. Bäckvall<br />

(Ed.), Modern Oxidation Methods, Wiley-VCH, Weinheim, 2004.<br />

2. (a) Zassinovich, G.; Mestroni, G.; Gladiali, S. Chem. Rev. 1992, 51,<br />

1051-1069; (b) Noyori, R.; Hashiguchi, S. Acc. Chem. Res. 1997, 30,<br />

97-102; (c) Gladiali, S.; Mestroni, G. in Beller, M.; Bolm, C. (Eds.),<br />

Transition Metals for Organic Synthesis, Wiley-VCH, Weinheim, 2 nd<br />

Ed. 2004, 145-166; (d) Blaser, H.-U.; Malan, C.; Pugin, B.; Spindler,<br />

F.; Studer, M. Adv. Synth. Catal. 2003, 345, 103-151; (e) Gladiali, S.;<br />

Alberico, E. Chem. Soc. Rev. 2006, 35, 226–236; (f) Samec, J. S. M.;<br />

Bäckvall, J.-E.; Andersson, P. G.; Brandt, P. Chem. Soc. Rev. 2006,<br />

35, 237–248.<br />

3. Selected recent examples of transfer hydrogenations: (a) Schlatter, A.;<br />

Kundu, M. K.; Woggon, W.-D. Angew. Chem. Int. Ed. 2004, 43,<br />

6731–6734; (b) Xue, D.; Chen, Y.-C.; Cui, X.; Wang, Q.-W.; Zhu, J.;<br />

Deng, J.-G. J. Org. Chem. 2005, 70, 3584–3591; (c) Wu, X.; Li, X.;<br />

King, F.; Xiao, J. Angew. Chem. Int. Ed. 2005, 44, 3407–3411; (d)<br />

Hayes, A. M.; Morris, D. J.; Clarkson, G. J.; Wills, M. J. Am. Chem.<br />

Soc. 2005, 127, 7318–7319; (e) Baratta, W.; Chelucci, G.; Gladiali,<br />

S.; Siega, K.; Toniutti, M.; Zanette, M.; Zangrando, E.; Rigo, P.<br />

Angew. Chem. Int. Ed. 2005, 44, 6214–6219; (f) Enthaler, S.;<br />

Jackstell, R.; Hagemann, B.; Junge, K.; Erre, G.; Beller, M. J.<br />

Organomet. Chem. 2006, 691, 4652–4659; (g) Enthaler, S.;<br />

Hagemann, B.; Bhor, S.; Anilkumar, G.; Tse, M. K.; Bitterlich, B.;<br />

Junge, K.; Erre, G.; Beller, M. Adv. Synth. Catal. 2007, accepted.<br />

4. (a) Cheung, F. K.; Hayes, A. M.; Hannedouche, J.; Yim, A. S. Y.;<br />

Wills, M. J. Org. Chem. 2005, 70, 3188-3197; (b) Matharu, D. S.;<br />

Morris, D. J.; Kawamoto, A. M.; Clarkson, G. J.; Wills, M. Org. Lett.<br />

2005, 7, 5489–5491; (c) Peach, P.; Cross, D. J.; Kenny, J. A.; Mann,<br />

I.; Houson, I.; Campbell, L.; Walsgrove, T.; Wills, M. Tetrahedron,<br />

2006, 62, 1864-1876; (c) Hamada, T.; Torii, T.; Izawa, K.; Ikariya, T.<br />

Tetrahedron 2004, 60, 7411-7417.<br />

5. The price for iron is relatively low (current price for 1 t of iron ~300<br />

US-dollar) in comparison to rhodium (~5975 US-dollar per troy<br />

ounce), ruthenium (~870 US-dollar per troy ounce) or iridium (~460<br />

US-dollar per troy ounce) due to the widespread abundance of iron in<br />

earth´s crust and the easy accessibility. Furthermore, iron is an<br />

essential trace element (daily dose for humans 5–28 mg) and plays an<br />

important role in an extensive number of biological processes.<br />

6. Bolm, C.; Legros, J.; Le Paih, J.; Zani, L. Chem. Rev. 2004, 104,<br />

6217–6254.<br />

7. Noyori, R.; Umeda, I.; Ishigami, T. J. Org. Chem. 1972, 37, 1542–<br />

1545.<br />

8. (a) Jothimony, K.; Vancheesan, S. J. Mol. Catal. 1989, 52, 301–304;<br />

(b) Jothimony, K.; Vancheesan, S.; Kuriacose, J. C. J. Mol. Catal.<br />

1985, 32, 11–16.<br />

9. Bianchini, C.; Farnetti, E.; Graziani, M.; Peruzzini, M.; Polo, A.<br />

Organometallics 1993, 12, 3753–3761.<br />

10. Chen, J.-S.; Chen, L.-L.; Xing, Y.; Chen, G.; Shen, W.-Y.; Dong, Z.-<br />

R.; Li, Y.-Y.; Gao, J.-X. Huaxue Xuebao 2004, 62, 1745–1750.<br />

11. (a) Enthaler, S.; Hagemann, B.; Erre, G.; Junge, K.; Beller M. Chem.<br />

Asian. J. 2006, 1, 598–604; (b) Enthaler, S.; Erre, G.; Tse, M. K.;<br />

Junge, K.; Beller, M. Tetrahedron Lett. 2006, 47, 8095–8099.<br />

12. After crystallization step Fe(III)citrate was subjected to ICP-AAS to<br />

ensure the non-existence of impurifications e.g. Rh, Ru or Ir.<br />

13. Adkins, H.; Elofson, R. M.; Rossow, A. G.; Robinson, C. C. J. Am.<br />

Chem. Soc. 1949, 71, 3622–3629.<br />

14. (a) Sun, C.-L.; Pang, R.-F.; Zhang, H.; Yang, M. Bioorg. Med. Chem.<br />

Lett. 2005, 15, 3257–3262; (b) Fabbri, C.; Bietti, M.; Lanzalunga, O.<br />

J. Org. Chem. 2005, 70, 2720–2728; (c) Kandanarachchi, P. H.;<br />

Autrey, T.; Franz, J. A. J. Org. Chem. 2002, 67, 7937–7945.<br />

15. Kadish, K. M.; Finikova, O. S.; Espinosa, E.; Gros, C. P.; De Stefano,<br />

G.; Cheprakov, A. V.; Beletskaya, I. P.; Guilard, R. Journal<br />

Porphyrins and Phthalocyanines 2004, 8, 1062–1066.<br />

16. In case of catalyst containing ligand 1a the formation of ironporphyrin<br />

complex was confirmed by MS.<br />

17. Alexiou, C.; Lever, A. B. P. Coord. Chem. Rev. 2001, 216–217, 45–<br />

54.<br />

18. Hodge, J. A.; Hill, M. G.; Gray, H. B. Inorg. Chem. 1995, 34, 809–<br />

812.<br />

19. After stirring FeCl2 and ligand 1a for 16 hours at 65 °C a sample was<br />

taken and immediately submitted to MS investigations. Iron porphyrin<br />

complex: calculated: m/z = 806 found: MS (EI): m/z = 806.<br />

20. (a) Hu, C.; Noll, B. C.; Schulz, C. E.; Scheidt, W. R. Inorg. Chem.<br />

2007, 46, 619–621; (b) Melin, F.; Choua, S.; Bernard, M.; Turek, P.;<br />

Weiss, J. Inorg. Chem. 2006, 45, 10750–10757; (c) Liu, C.; Shen, D.-<br />

M.; Chen, Q.-Y. Eur. J. Org. Chem. 2006, 12, 2703–2706; (d) Ruzie,<br />

C.; Even, P.; Ricard, D.; Roisnel, T.; Boitrel, B. Inorg. Chem. 2006,<br />

45, 1338–1348; (e) Even, P.; Ruzie, C.; Ricard, D.; Boitrel, B. Org.<br />

Lett. 2005, 7, 4325–4328.<br />

21. (a) Lindsey, J. S.; Hsu, H. C.; Schreiman, I. C.; Tetrahedron Lett.<br />

1986, 27, 4969–4970; (b) Lindsey, J. S.; Schreiman, I. C.; Hsu, H. C.;<br />

Kearney, P. C.; Marguerettaz, A. M. J. Org. Chem. 1987, 52, 827–<br />

836.


6. Transfer hydrogenation<br />

6.2. Transfer hydrogenation of prochiral ketones with ruthenium catalysts<br />

6.2.1. Efficient transfer hydrogenation of ketones in the presence of<br />

ruthenium N–heterocyclic carbene catalysts<br />

Stephan Enthaler, Ralf Jackstell, Bernhard Hagemann, Kathrin Junge, Giulia Erre and<br />

Matthias Beller*, J. Organomet. Chem. 2006, 691, 4652–4659.<br />

Contributions<br />

Imidazolium salts 1, 2, 5, 6, 8 and 9 were synthesized by Dr. R. Jackstell and Imidazolium<br />

salts 11 and 12 were a chemical gift by solvent innovation.<br />

125


Abstract<br />

Efficient transfer hydrogenation of ketones in the presence<br />

of ruthenium N-heterocyclic carbene catalysts<br />

Stephan Enthaler, Ralf Jackstell, Bernhard Hagemann, Kathrin Junge,<br />

Giulia Erre, Matthias Beller *<br />

Leibniz-Institut für Katalyse e.V. an der <strong>Universität</strong> <strong>Rostock</strong>, Albert-Einstein-Strasse 29a, 18059 <strong>Rostock</strong>, Germany<br />

Received 27 April 2006; received in revised form 7 July 2006; accepted 13 July 2006<br />

Available online 27 July 2006<br />

Novel ruthenium carbene complexes have been in situ generated and tested for the transfer hydrogenation of ketones. Applying<br />

Ru(cod)(methylallyl)2 in the presence of imidazolium salts in 2-propanol and sodium-2-propanolate as base, turnover frequencies up<br />

to 346 h 1 have been obtained for reduction of acetophenone. A comparative study involving ruthenium carbene and ruthenium phosphine<br />

complexes demonstrated the higher activity of ruthenium carbene complexes.<br />

Ó 2006 Elsevier B.V. All rights reserved.<br />

Keywords: Ruthenium; Homogeneous catalysis; Transfer hydrogenation; N-heterocyclic carbenes; Ketones<br />

1. Introduction<br />

Journal of Organometallic Chemistry 691 (2006) 4652–4659<br />

The preparation of alcohols has become an important<br />

field of activity for transition metal catalyzed reactions<br />

[1]. Within the different catalytic approaches developed,<br />

for instance addition of organometallic reagents to carbonyl<br />

compounds, hydroxylation of olefins, functionalization<br />

reactions of epoxides, the hydrogenation of ketones<br />

or aldehydes is the most powerful tool with respect to<br />

industrial applications. In particular, transfer hydrogenations<br />

represent a potent strategy, because of high atom efficiency,<br />

no need of pressure, and economic as well as<br />

environmental advantages [2]. In more detail, a broad<br />

scope of alcohols is accessible by transfer hydrogenation<br />

using non-toxic hydrogen donors under mild reaction conditions<br />

in the presence of various metal catalysts, such as<br />

Ir, Rh or Ru [2d]. Noteworthy, a prerequisite for achieving<br />

high activity and selectivity is the fine tuning of the metal<br />

catalyst by introduction of ligands. So far the development<br />

of new ligands for catalytic reductions focused predominantly<br />

on phosphines and amines.<br />

* Corresponding author. Tel.: +49 381 1281 0; fax: +49 381 1281 5000.<br />

E-mail address: matthias.beller@ifok-rostock.de (M. Beller).<br />

0022-328X/$ - see front matter Ó 2006 Elsevier B.V. All rights reserved.<br />

doi:10.1016/j.jorganchem.2006.07.013<br />

www.elsevier.com/locate/jorganchem<br />

More recently carbene ligands found increasing interest<br />

for exploiting new catalytic reactions [3]. Stable N-heterocyclic<br />

carbenes (NHC) were first introduced in the early<br />

1990’s by Arduengo et al. [4]. Since the mid 1990’s Herrmann<br />

et al. [5] and then the groups of Bertrand [6], Blechert<br />

[7], Cavell [8], Fürstner [9]. Glorius [10], Grubbs [11],<br />

Nolan [12] and others [13] demonstrated the catalytic<br />

potential of NHC metal complexes. In this context we<br />

reported that palladium carbene complexes are excellent<br />

catalysts for different coupling reactions of aryl halides<br />

and telomerizations [14].<br />

With regard to transfer hydrogenations different carbene<br />

or carbene-phosphine-systems containing Rh [15], Ir<br />

[15,16], Ru[17] and Ni [18] have been reported. Excellent<br />

turnover frequencies up to 120,000 h 1 were reported by<br />

the groups of Baratta and Herrmann applying a ruthenium-carbene-phosphine-catalysts<br />

[19]. However, for<br />

reduction of a typical substrate, e.g. acetophenone, with<br />

phosphine-free ruthenium-carbene catalysts lower turnover<br />

frequencies (TOF 333 h 1 ) [17e] were achieved in comparison<br />

to iridium (500 h 1 ) [16c] and rhodium systems<br />

(583 h 1 ) [15b]. Due to the economical benefit of ruthenium<br />

metal compared to rhodium or iridium and the<br />

advantages of phosphine-free systems, it is an important


goal to search for more active ruthenium carbene catalysts.<br />

Herein, we report the application of novel in situ prepared<br />

ruthenium carbene catalysts in the reduction of different<br />

ketones.<br />

2. Results and discussion<br />

S. Enthaler et al. / Journal of Organometallic Chemistry 691 (2006) 4652–4659 4653<br />

Based on our experience in the synthesis of carbene<br />

ligands and their application in homogeneous catalysis,<br />

we became interested in demonstrating the usefulness of<br />

carbene-complexes in ruthenium-catalyzed transfer hydrogenations<br />

[20]. From a practical point of view the application<br />

of in situ prepared catalysts has significant advantages.<br />

Thus, we used a small library of various imidazolium salts<br />

(Scheme 1, 1–12) as carbene precursors. In exploratory<br />

experiments, 2-propanol-based transfer hydrogenation of<br />

acetophenone was examined. In order to ensure complete<br />

formation of the active catalyst a 2-propanol solution of<br />

1 mol% ruthenium-source and 1 mol% 1,3-bis(2,6-di-i-propylphenyl)-imidazolium<br />

chloride (1) is stirred in the presence<br />

of 5 mol% sodium 2-propylate for 16 h at 65 °C.<br />

Initial investigations showed a crucial effect on reactivity<br />

by using different ruthenium sources such as [RuCl 2-<br />

(C 6H 6)] 2,Ru 3(CO) 12, and Ru(cod)(methylallyl) 2 in combination<br />

with imidazolium salt 1 (Table 1). Best conversion<br />

and yield are obtained for Ru(cod)(methylallyl)2/1 at<br />

100 °C (Table 1, entry 8). Noteworthy, there is a significant<br />

temperature effect on the reaction rate (Table 1, entries 5–8).<br />

N +<br />

N<br />

N +<br />

N<br />

N +<br />

Cl -<br />

Br -<br />

N +<br />

N<br />

N +<br />

N<br />

N +<br />

It is well-known that transfer hydrogenations are sensitive<br />

to the nature of the base. Thus, the influence of sodium<br />

2-propylate, potassium tert-butylate, potassium carbonate<br />

and sodium hydroxide on selectivity and conversion was<br />

investigated. In all cases excellent selectivity (>99%) is<br />

N +<br />

- -<br />

BF4 BF4 N<br />

1 2 3 4<br />

Cl -<br />

5 6 7 8<br />

Cl<br />

N<br />

(CH2) 2<br />

Cl<br />

N<br />

N<br />

N<br />

- Cl- Br- Br- 9 10 11 12<br />

N +<br />

N<br />

N +<br />

Cl<br />

Cl -<br />

Scheme 1. Selection of imidazolium salts.<br />

Table 1<br />

Screening of various ruthenium sources and yield-temperature dependency<br />

for the transfer hydrogenation of acetophenone (13) a<br />

O<br />

[Ru]/(1), Na-2-OPr<br />

2-PrOH, 1h<br />

N +<br />

N<br />

N +<br />

N<br />

N +<br />

Cl -<br />

Tos -<br />

OH<br />

13 13a<br />

Entry Source Temperature (°C) Yield (%) b<br />

1 [RuCl2(C6H6)] 2 90 9<br />

2 [RuCl2(C6H6)]2 100 80<br />

3 Ru3(CO)12 90 40<br />

4 Ru3(CO)12 100 16<br />

5 Ru(cod)(methylallyl) 2 70 2<br />

6 Ru(cod)(methylallyl)2 80 43<br />

7 Ru(cod)(methylallyl)2 90 55<br />

8 Ru(cod)(methylallyl) 2 100 >99<br />

a 6<br />

Reaction conditions: in situ catalyst: 1.3 · 10 mol Ru3(CO) 12,<br />

1.9 · 10 6 mol [RuCl2(C6H6)] 2 or 3.8 · 10 6 mol Ru(cod)(methylallyl) 2,<br />

3.8 · 10 6 mol imidazolium salt 1 and 1.9 · 10 5 mol Na–2-OPr in 2.0 mL<br />

2-propanol for 16 h at 65 °C, addition of 3.8 · 10 4 mol acetophenone<br />

(13), reaction 1 h at described temperature.<br />

b<br />

Conversion is determined by GC analysis (50 m Lipodex E, 95–200 °C)<br />

with diglyme as internal standard.


4654 S. Enthaler et al. / Journal of Organometallic Chemistry 691 (2006) 4652–4659<br />

observed. Best conversion after 1 h at 100 °C are obtained<br />

in the presence of sodium 2-propylate (>99%) and potassium<br />

tert-butylate (95%). However, poor yields of 1-phenylethanol<br />

are achieved with potassium carbonate (53%).<br />

Interestingly, sodium hydroxide, one of the most common<br />

bases for transfer hydrogenation, induced only low conversion<br />

(62%). By increasing the amount of base a further<br />

acceleration of reaction rate is recorded, while no reaction<br />

occurred in the absence of base.<br />

Next, the influence of the ligand concentration was<br />

investigated by variation of the metal to ligand-ratio. When<br />

increasing the equivalents of ligand per metal a negative<br />

effect on the reaction rate is observed (Table 2). We assume<br />

a catalyst deactivation by more than one carbene ligand,<br />

due to suppressing the metal hydride formation or blocking<br />

the active binding site for the substrate. However, 1 equiv.<br />

Table 2<br />

Influence of metal-ligand-ratio on the reduction of acetophenone (13)<br />

O<br />

Ru(cod)methylallyl 2/(1), Na-2-OPr<br />

2-PrOH, 100 ˚C<br />

of ligand is necessary for achieving good conversion.<br />

Hence, transfer hydrogenations in absence of the carbene<br />

gave only moderate yield (Table 2, entry 1).<br />

The stability of metal carbene complexes against moisture<br />

and oxygen has been documented [13]. Thus, the addition<br />

of water (10 mol%) to the reaction mixture decreased<br />

only slightly the conversion to 73% (TOF: 73 h 1 ). Even<br />

in the presence of 100 mol% of water the catalyst showed<br />

significant activity (TOF: 54 h 1 ).<br />

To classify the potential of our catalytic system we compared<br />

Ru(cod)(methylallyl)2/1 with Ru(cod)(methylallyl)2/<br />

PPh3 and Ru(cod)(methylallyl)2/PCy3 (Fig. 1). More<br />

specifically, we studied the behaviour of Ru(cod)(methylallyl)2/1<br />

and Ru(cod)(methylallyl)2/PPh3 by monitoring the<br />

conversion at different reaction times. The results showed<br />

similar catalytic behaviour at the beginning of the reaction.<br />

OH<br />

13 13a<br />

Entry Carbene:metal Substrate:metal Base:metal Time (h) Yield (%) c<br />

1 a<br />

2 a<br />

3 a<br />

4 a<br />

5 b<br />

6 b<br />

7 b<br />

TOF (h 1 ) d<br />

0 100 5 1 41 41<br />

1 100 5 1 >99 99<br />

2 100 5 1 61 61<br />

10 100 5 1 57 57<br />

1 5000 100 12 83 346<br />

2 5000 100 12 81 338<br />

10 5000 100 12 67 279<br />

a 6<br />

Reaction conditions: in situ catalyst: 3.8 · 10 mol Ru(cod)(methylallyl)2, 3.8 · 10 6 mol imidazolium salt 1 and 1.9 · 10 5 mol Na–2-OPr in 2.0 mL<br />

2-propanol for 16 h at 65 °C, addition of 3.8 · 10 4 mol acetophenone (13), reaction temperature 100 °C.<br />

b 7<br />

Reaction conditions: in situ catalyst: 9.7 · 10 mol Ru(cod)(methylallyl)2, 9.7 · 10 7 mol imidazolium salt 1 and 9.7 · 10 5 mol Na–2-OPr in 5.0 mL<br />

2-propanol for 16 h at 65 °C, addition of 4.85 · 10 3 mol acetophenone (13), reaction temperature 100 °C.<br />

c<br />

Conversion is determined by GC analysis (50 m Lipodex E, 95–200 °C) with diglyme as internal standard.<br />

d Turnover frequency = mol product/(mol catalyst · time), determined after 12 h.<br />

yield (%)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

O<br />

Carben (1)<br />

PPh3<br />

PCy3<br />

Ru(cod)(methylallyl) 2 /(1) or PPh 3,<br />

Na-2-OPr<br />

OH<br />

2-PrOH, 100 ˚C<br />

13 13a<br />

0 2 4 6 8 10 12<br />

time (h)<br />

Fig. 1. Comparative study using imidazolium salt 1 and PPh3 as ligands. Note: Reaction conditions: in situ catalyst: 9.7 · 10 7 mol Ru(cod)(methylallyl)2,<br />

9.7 · 10 7 mol imidazolium salt 1 or PPh3 and 9.7 · 10 5 mol Na-2-OPr in 5.0 mL 2-propanol for 16 h at 65 °C, addition of 4.85 · 10 3 mol acetophenone<br />

(13), reaction at 100 °C. Conversion is determined by GC analysis (50 m Lipodex E, 95–200 °C) with diglyme as internal standard.


However, during the reaction a higher deactivation rate of<br />

the PPh3-system is detected, which resulted in a lower yield<br />

of 1-phenylethanol after 12 h (68% vs 83%). The<br />

Ru(cod)(methylallyl)2/PCy3 yielded comparable amounts<br />

of 1-phenylethanol to Ru(cod)(methylallyl)2/1.<br />

As shown in Table 3 we examined 12 examples out of<br />

the growing number of carbene precursors (1–12) under<br />

the previously optimized conditions for the transfer hydrogenation<br />

of acetophenone (13). In order to estimate differences<br />

between the various carbene precursors we applied<br />

low catalyst loadings (0.02 mol%) at 100 °C. After 12 h<br />

average turnover frequencies up to 346 h 1 are achieved<br />

for the preparation of 1-phenylethanol applying ligand 1.<br />

Summarizing the activities of 4,5-dihydroimidazolium<br />

salts, no pronounced influence is observed by variation of<br />

substitutents at the nitrogen atoms (2,6-di-iso-propylphenyl<br />

or mesitylene groups) or by changing the anion of the<br />

imidazolium salt (Table 3, entries 2–4). On the other hand<br />

by introduction of methyl groups in the 4,5-position of the<br />

imidazolium unit a depletion of activity is monitored<br />

(Table 3, entries 5 and 6).<br />

In general, application of N-alkyl carbenes led to a<br />

lower activity compared to N-aryl carbenes (Scheme 1, 8–<br />

12). In the presence of 1-ethyl-3-methylimidazolium bromide<br />

([EMIM]Br, 11) and 1-butyl-3-methylimidazolium<br />

bromide ([BMIM]Br, 12), which are usually used as ionic<br />

liquids [21], the recorded yields were lower (Table 3, entries<br />

11 and 12).<br />

Table 3<br />

Variation of imidazolium salts in the transfer hydrogenation of acetophenone<br />

(13) a<br />

O<br />

Ru(cod)(methylallyl) 2/(1-12), Na-2-OPr<br />

2-PrOH, 100 ˚C, 12h<br />

OH<br />

13 13a<br />

Entry Imidazolium salt Yield (%) b<br />

S. Enthaler et al. / Journal of Organometallic Chemistry 691 (2006) 4652–4659 4655<br />

TOF (h 1 ) c<br />

TON d<br />

1 1 83 346 4150<br />

2 2 75 313 3750<br />

3 3 68 283 3400<br />

4 4 73 304 3650<br />

5 5 53 221 2650<br />

6 6 61 254 3050<br />

7 7 62 258 3100<br />

8 8 67 279 3350<br />

9 9 77 320 3850<br />

10 10 69 288 3450<br />

11 11 54 225 2700<br />

12 12 38 158 1900<br />

a 7<br />

Reaction conditions: in situ catalyst: 9.7 · 10 mol Ru(cod)(methylallyl)2,<br />

9.7 · 10 7 mol imidazolium salt and 9.7 · 10 5 mol Na-2-OPr in<br />

5.0 mL 2-propanol for 16 h at 65 °C, addition of 4.85 · 10 3 mol acetophenone<br />

(13), reaction for 12 h at 100 °C.<br />

b<br />

Conversion is determined by GC analysis (50 m Lipodex E, 95–200 °C)<br />

with diglyme as internal standard.<br />

c<br />

Turnover frequency = mol product/(mol catalyst · time), determined<br />

after 12 h.<br />

d<br />

Turnover number = mol product/mol catalyst, determined after 12 h.<br />

In order to demonstrate the usefulness of the catalysts in<br />

a more general manner we employed the catalyst system<br />

Ru(cod)(methylallyl)2/1 in the transfer hydrogenation of<br />

nine aromatic and aliphatic ketones (Table 4).<br />

In general, all substrates were hydrogenated with excellent<br />

chemoselectivity (>99%). Best activity (TOF up to<br />

338 h 1 ) is achieved with dialkyl ketones (Table 4, entries<br />

Table 4<br />

Scope and limitations of Ru(cod)(methylallyl)2/1-system-catalyzed ketone<br />

reduction a<br />

O<br />

OH<br />

Ru(cod)(methylallyl) 2/(1), Na-2-OPr<br />

R 1<br />

R 2<br />

14-22<br />

2-PrOH, 100 ˚C, 12h<br />

R 1<br />

R 2<br />

14a-22a<br />

Entry Compound Ketone Conversion (%) b TOF (h 1 ) c<br />

1 14 O<br />

2 15<br />

3 16<br />

4 17<br />

5 18<br />

6 19<br />

7 20<br />

8 21<br />

9 22<br />

Cl<br />

MeO<br />

H 3C<br />

O<br />

O<br />

OMe<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

68 [96] 285<br />

41 171<br />

62 [92] 258<br />

68 [97] 283<br />

67 [93] 281<br />

Cl 2 6<br />

40 166<br />

77 321<br />

81 338<br />

a 7<br />

Reaction conditions: in situ catalyst: 9.7 · 10 mol Ru(cod)(methylallyl)2,<br />

9.7 · 10 7 mol imidazolium salt 1 and 9.7 · 10 5 mol Na-2-OPr in<br />

5.0 mL 2-propanol for 16 h at 65 °C, addition of 4.85 · 10 3 mol ketone,<br />

reaction for 12 h at 100 °C.<br />

b<br />

Conversion is determined by GC analysis (14 (25 m Lipodex E,<br />

100 °C), 15 (50 m Lipodex E, 90–105 °C), 16 (25 m Lipodex E, 80–180 °C),<br />

17 (30 m HP Agilent Technologies 50–300 °C), 18 (25 m Lipodex E, 90–<br />

180 °C), 19 (50 m Lipodex E, 90–180 °C), 20–22 (30 m HP Agilent Technologies<br />

50–300 °C)) with diglyme as internal standard. In brackets the<br />

conversion after 24 h is given.<br />

c<br />

Turnover frequency = mol product/(mol catalyst · time), determined<br />

after 12 h.


4656 S. Enthaler et al. / Journal of Organometallic Chemistry 691 (2006) 4652–4659<br />

8 and 9). In comparison, para-substituted acetophenones<br />

containing an electron-withdrawing group (Table 4, entry<br />

1) showed better conversion than para-substituted substrates<br />

with an electron-donating group (Table 4, entry<br />

2). Noteworthy, by changing the electron-donating group<br />

from para- toortho-position a significant increase of the<br />

yield is detected (Table 4, entries 2 and 4). We assume<br />

for 17 a possible second coordination site at the metal center.<br />

No major change in activity is observed for substitution<br />

adjacent to the carbonyl group by an ethyl group,<br />

whereas introduction of a chloromethyl deactivated the<br />

catalyst (Table 4, entries 5 and 6). Moderate activity is<br />

monitored when increasing the bulkiness next to the active<br />

center by a cyclopropyl group (Table 4, entry 7).<br />

Finally, we were interested in mechanistic aspects. In<br />

general, for transition metal catalyzed transfer hydrogenation<br />

two mechanisms are accepted, designated as direct<br />

hydrogen transfer, via formation of a six-membered cyclic<br />

transition state composed of donor, metal and acceptor,<br />

and the hydridic route which shows two possible pathways,<br />

the monohydride or dihydride mechanism. In more detail,<br />

formation of a monohydride-metal-complex promoted an<br />

exclusive hydride transfer from carbon (donor) to carbonyl<br />

carbon (acceptor) (Scheme 2, pathway A), whereas a<br />

hydride transfer from carbon (donor) to carbonyl carbon<br />

(acceptor) as well as to the carbonyl oxygen (acceptor)<br />

was proposed for a dihydride-metal-complex formation<br />

(Scheme 2, pathway B). Indications for both pathways<br />

were published by Bäckvall et al. and other groups, when<br />

following the hydride transfer catalyzed by various metal<br />

complexes [22].<br />

Mechanistic studies have been mostly published for catalysts<br />

containing phosphines, amines or cyclopentadienyls<br />

as ligands [22a]. For transition metal complexes containing<br />

carbene ligands Faller and Crabtree described investigations<br />

on an iridium dicarbene system [16c]. They assumed<br />

O<br />

[M-H*]<br />

O<br />

D<br />

D<br />

H<br />

O<br />

*H<br />

pathway A pathway B<br />

OH<br />

O H<br />

O<br />

-acetone H*<br />

*H<br />

-acetone<br />

Ru(cod)(methylallyl) 2/1,<br />

Na-2-OPr-d 7<br />

2-PrOH-d 8, 100 ˚C<br />

O<br />

a monohydride mechanism, because the hydride is mainly<br />

transferred in the 1-position of acetophenone. So far there<br />

is no mechanistic investigation known in transfer hydrogenations<br />

applying Ru carbene catalysts.<br />

Reaction of ketone 20 (‘‘radical clock’’-substrate) with<br />

2-propanol in the presence of 1 mol% Ru(cod)(methylallyl)<br />

2/1 gave only the corresponding cyclopropyl phenyl<br />

alcohol (>99% by 1 H NMR). Apparently, there is no<br />

radical induced reduction [23]. Owing to this a radical<br />

reduction mechanism promoted by sodium alkoxides can<br />

be also excluded, whereby the transition metal plays only<br />

a marginal role [24]. This assumption is also confirmed<br />

by performing the reduction of acetophenone (13) in the<br />

presence of base and in the absence of ruthenium catalyst.<br />

Here, no reduction product was detected.<br />

Next, we followed the transfer of hydrogen from the<br />

donor molecule into the product by applying a deuterated<br />

donor [25]. The catalytic precursor is generated by stirring<br />

a solution of 2-propanol-d8, Ru(cod)(methylallyl)2 and imidazolium<br />

salt 1 in the presence of sodium 2-propanolate-d7<br />

for 16 h at 65 °C. Then, acetophenone (13) was added and<br />

the solution was stirred for 30 min at 100 °C. As main<br />

product (>99%) 23 was observed by 1 H NMR (Scheme<br />

3) [26]. The result showed an exclusive transfer of the deuterium<br />

into the carbonyl group, so that no C–H activation<br />

on the substrate occurred under the described conditions.<br />

Furthermore, this result rules out enol formation in the catalytic<br />

cycle [27].<br />

To clarify the transfer of hydrogen from the hydrogen<br />

donor into the substrate the reaction was run with 2-propanol-d<br />

1 (hydroxy-group deuterated) as solvent/donor and<br />

sodium 2-propylate as base. In the transfer hydrogenation<br />

of acetophenone (13) we obtained a mixture of two different<br />

deuterated 1-phenylethanols (Scheme 2, 24a and 24b).<br />

Here, a scrambling of the transferred proton and deuteride<br />

is found (24a and 24b = 1:1). In conclusion the non-specific<br />

[H-M-H*]<br />

Scheme 2. Comparison of monohydride and dihydride mechanism for transfer hydrogenations.<br />

Ru(cod)(methylallyl) 2/1,<br />

Na-2-OPr<br />

2-PrOD, 100 ˚C<br />

23 13<br />

24a 24b<br />

>99% 50% 50%<br />

Scheme 3. Deuterium incorporation into acetophenone catalyzed by Ru(cod)(methylallyl) 2/1-system in the presence of base.<br />

O<br />

H<br />

D<br />

OH<br />

H*<br />

+<br />

O<br />

D<br />

OH*<br />

H<br />

50% 50%<br />

H


migration is in agreement with the dihydride mechanism,<br />

implying a formation of metal dihydride species in the catalytic<br />

cycle [2d].<br />

3. Summary<br />

We demonstrated the successful application of in situ<br />

prepared ruthenium catalysts containing carbene ligands<br />

in the transfer hydrogenation of various ketones. In the<br />

reduction of acetophenone (13) turnover frequencies up<br />

to 346 h 1 were found for a catalyst system containing<br />

Ru(cod)(methylallyl)2/1,3-bis(2,6-di-i-propylphenyl)-imidazolium<br />

chloride (1). Mechanistic experiments indicated<br />

the transfer of hydrogen from the donor molecule into<br />

the substrate via a dihydride mechanism.<br />

4. Experimental section<br />

4.1. General<br />

All manipulations were performed under argon atmosphere<br />

using standard Schlenk techniques. Unless specified,<br />

all chemicals are commercially available and used as<br />

received. Sodium 2-propylate was prepared by reacting<br />

sodium with 2-propanol under an argon atmosphere. 2-<br />

Propanol was used without further purification (purchased<br />

from Fluka, dried over molecular sieves). Imidazolium<br />

salts 1, 2, 5, 6, 8 and 9 were synthesized according to the<br />

published protocols [4,28]. Imidazolium salts 11 and 12<br />

were a gift by Solvent Innovation. Imidazolium salts 3, 4,<br />

7 and 10 are commercially available by Strem. All ketones<br />

were dried over CaH 2, distilled in vacuum and stored under<br />

argon, except ketones 17 and 19, which were cycled with<br />

vacuum-argon and stored under argon.<br />

4.2. General procedure for catalytic transfer hydrogenation<br />

of ketones<br />

In a 10 mL Schlenk tube, the in situ catalyst (9.7 ·<br />

10 7 mol) was prepared by stirring a solution of<br />

Ru(cod)(methylallyl) 2 (9.7 · 10 7 mol), imidazolium salt<br />

(9.7 · 10 7 mol) and sodium 2-propylate (4.85 · 10 6 mol)<br />

in 1.0 mL 2-propanol for 16 h at 65 °C. After addition of<br />

the corresponding ketone (4.85 · 10 3 mol) and the internal<br />

standard diglyme in 4.0 mL 2-propanol the Schlenk tube<br />

was sealed and the reaction mixture was heated to 100 °C.<br />

After 12 h the conversion was measured by GC without further<br />

purification. In the case of 1 H NMR determination of<br />

the yield, the solvent was removed in vacuum and the residue<br />

was dissolved in CDCl3 and submitted to 1 H NMR.<br />

4.3. Procedure for transfer hydrogenation of acetophenone<br />

with 2-propanol-d8 as hydride source<br />

S. Enthaler et al. / Journal of Organometallic Chemistry 691 (2006) 4652–4659 4657<br />

In a 10 mL Schlenk tube, Ru(cod)(methylallyl) 2 (3.8 ·<br />

10 6 mol), imidazolium salt 1 (3.8 · 10 6 mol) and sodium<br />

2-propylate-d 7 (1.9 · 10 5 mol, prepared by reacting<br />

sodium with 2-propanol-d8) was solved in 1.0 mL 2-propanol-d8<br />

and stirred for 16 h at 65 °C. After addition of the<br />

acetophenone (13) (3.8 · 10 4 mol) in 2.0 mL 2-propanold8<br />

the reaction mixture was heated to 100 °C for 30 min.<br />

The solution was cooled to r.t. and filtrated over a plug<br />

of silica. The conversion was determined by 1 H NMR.<br />

4.4. Procedure for transfer hydrogenation of acetophenone<br />

with 2-propanol-d1 as hydride source<br />

In a 10 mL Schlenk tube, Ru(cod)(methylallyl)2 (3.8 ·<br />

10 6 mol), imidazolium salt 1 (3.8 · 10 6 mol) and sodium<br />

2-propylate (1.9 · 10 5 mol, prepared by reacting sodium<br />

with 2-propanol) was solved in 1.0 mL 2-propanol-d 1 (deuterium<br />

fixed as hydroxyl proton) and stirred for 16 h at<br />

65 °C. The reaction mixture was heated to 100 °C for<br />

10 min after addition of the acetophenone (13) (3.8 ·<br />

10 4 mol) in 2.0 mL 2-propanol-d1. (To avoid side effects<br />

reaction was not run to <strong>full</strong> conversion.) The solution<br />

was cooled to r.t. and filtrated over a plug of silica. The solvent<br />

was removed in vacuum and the residue was solved in<br />

CDCl 3. The conversion was determined by 1 H NMR.<br />

4.5. Product characterization<br />

The obtained alcohols are known compounds. They<br />

were characterized by comparison with authentic samples<br />

and mass spectroscopy (Agilent Technologies 6890N,<br />

MSD 5973) or 1 H NMR (Bruker ARX-400). Product<br />

13a: m/z (%) = 122 (M + , 21); 107 (72); 79 (100); 51 (34);<br />

43 (36); 39 (13); 32 (15). Product 14a: m/z (%) = 156<br />

(M + , 26); 141 (100); 121 (14); 113 (33); 103 (9); 77 (85);<br />

51 (12); 43 (22). Product 15a: m/z (%) = 152 (M + , 24);<br />

137 (100); 134 (32); 119 (20); 109 (48); 94 (38); 91 (34); 77<br />

(46); 65 (31); 51 (16); 43 (37); 39 (21). Product 16a: m/z<br />

(%) = 136 (M + , 32); 121 (100); 117 (14); 91 (94); 77 (53);<br />

65 (29); 51 (16); 43 (57); 39 (23). Product 17a: m/z<br />

(%) = 152 (M + , 31); 137 (100); 134 (20); 119 (10); 109<br />

(46); 94 (29); 91 (16); 77 (28); 65 (12); 43 (13). Product<br />

18a: m/z (%) = 152 (M + , 11); 107 (100); 79 (79); 51 (13).<br />

Product 19a: m/z (%) = 156 (M + , 3); 107 (100); 91 (7); 79<br />

(67); 77 (50); 51 (19). Product 20a: 1 H NMR (400 MHz,<br />

CDCl3) d (ppm) = 0.34 (1H, m); 0.44 (1H, m); 0.51 (1H,<br />

m); 0.58 (1H, m); 1.00 (1H, m); 2.31 (1H, s); 3.98 (1H, d,<br />

J = 8.16 Hz); 7.24 (3H, m); 7.99 (1H, d, J = 7.52 Hz).<br />

Product 21a: m/z (%) = 128 (M + , 17); 110 (34); 95 (18);<br />

81 (100); 67 (71); 55 (78); 41 (46). Product 22a: m/z<br />

(%) = 102 (M + , 3); 87 (26); 57 (100); 45 (83); 41 (53).<br />

Acknowledgements<br />

This work has been financed by the State of Mecklenburg-Pomerania<br />

and the Bundesministerium für Bildung<br />

und Forschung (BMBF). We thank Mrs. C. Voss, Mrs.<br />

C. Mewes, Mrs. M. Heyken, Mrs. S. Buchholz and Dr.<br />

C. Fischer (all Leibniz-Institut für Katalyse e.V.) for their


4658 S. Enthaler et al. / Journal of Organometallic Chemistry 691 (2006) 4652–4659<br />

excellent technical and analytical support. Solvent Innovation<br />

is grate<strong>full</strong>y thanked for a chemical gift.<br />

References<br />

[1] (a) M. Beller, C. Bolm (Eds.), Transition Metals for Organic<br />

Synthesis, second ed., Wiley–VCH, Weinheim, 2004;<br />

(b) B. Cornils, W.A. Herrmann, Applied Homogeneous Catalysis<br />

with Organometallic Compounds, second ed., Wiley–VCH, Weinheim,<br />

2002;<br />

(c) M. Kitamura, R. Noyori, in: S.-I. Murahashi (Ed.), Ruthenium in<br />

Organic Synthesis, Wiley–VCH, Weinheim, 2004;<br />

(d) E.N. Jacobsen, A. Pfaltz, H. Yamamoto (Eds.), Comprehensive<br />

Asymmetric Catalysis, Springer, Berlin, 1999;<br />

(e) R. Noyori, Asymmetric Catalysis in Organic Synthesis, Wiley,<br />

New York, 1994.<br />

[2] (a) S. Gladiali, E. Alberico, Chem. Soc. Rev. 35 (2006) 226–236;<br />

(b) G. Zassinovich, G. Mestroni, S. Gladiali, Chem. Rev. 51 (1992)<br />

1051–1069;<br />

(c) R. Noyori, S. Hashiguchi, Acc. Chem. Res. 30 (1997) 97–102;<br />

(d) S. Gladiali, G. Mestroni, in: M. Beller, C. Bolm (Eds.), Transition<br />

Metals for Organic Synthesis, second ed., Wiley–VCH, Weinheim,<br />

2004, pp. 145–166;<br />

(e) H.-U. Blaser, C. Malan, B. Pugin, F. Spindler, M. Studer, Adv.<br />

Synth. Catal. 345 (2003) 103–151.<br />

[3] (a) For reviews see: E. Peris, R.H. Crabtree, Coordin. Chem. Rev.<br />

248 (2004) 2239–2246;<br />

(b) M.C. Perry, K. Burgess, Tetrahedron: Asymmetr. 14 (2003) 951–<br />

961;<br />

(c) W.A. Herrmann, Angew. Chem. 114 (2002) 1342–1363;<br />

(d) T.M. Trnka, R.H. Grubbs, Acc. Chem. Res. 34 (2001) 18–29;<br />

(e) D. Bourisou, O. Guerret, F.P. Gabbai, G. Bertrand, Chem. Rev.<br />

100 (2000) 39–91;<br />

(f) W.A. Herrmann, C. Köcher, Angew. Chem. 109 (1997) 2256–<br />

2282.<br />

[4] (a) A.J. Arduengo III, R.L. Harlow, M. Kline, J. Am. Chem. Soc.<br />

113 (1991) 361–363;<br />

(b) A.J. Arduengo III, Acc. Chem. Res. 32 (1999) 913–921;<br />

(c) A.J. Arduengo III, R. Krafczyk, R. Schmutzler, H.A. Craig, J.R.<br />

Goerlich, W.J. Marshall, M. Unverzagt, Tetrahedron 55 (1999)<br />

14523–14534.<br />

[5] (a) C.W.K. Gstöttmayr, V.P.W. Böhm, E. Herdtweck, M.<br />

Grosche, W.A. Herrmann, Angew. Chem., Int. Ed. 41 (2002)<br />

1363–1365;<br />

(b) W.A. Herrmann, L.J. Gooßen, C. Köcher, G.R.J.. Artus,<br />

Angew. Chem., Int. Ed. 35 (1996) 2805–2807;<br />

(c) W.A. Herrmann, M. Elison, J. Fischer, C. Köcher, G.R.J.<br />

Artus, Angew. Chem., Int. Ed. 34 (1995) 2371–2374.<br />

[6] (a) V. Lavallo, Y. Canac, A. DeHope, B. Donnadieu, G. Bertrand,<br />

Angew. Chem., Int. Ed. 44 (2005) 7236–7239;<br />

(b) G. Bertrand, E. Diez-Barra, J. Fernandez-Baeza, H. Gornitzka,<br />

A. Moreno, A. Otero, R.I. Rodriguez-Curiel, J. Tejeda, Eur. J.<br />

Inorg. Chem. 11 (1999) 1965–1971.<br />

[7] (a) K. Vehlow, S. Maechling, S. Blechert, Organometallics 25 (2006)<br />

25–28;<br />

(b) R. Stragies, U. Voigtmann, S. Blechert, Tetrahedron Lett. 41<br />

(2000) 5465–5468;<br />

(c) S. Gessler, S. Randl, S. Blechert, Tetrahedron Lett. 41 (2000)<br />

9973–9976.<br />

[8] (a) J.W. Sprengers, J. Wassenaar, N.D. Clement, K.J. Cavell, Angew.<br />

Chem., Int. Ed. 44 (2005) 2026–2029;<br />

(b) N.D. Clement, K.J. Cavell, Angew. Chem., Int. Ed. 43 (2004)<br />

3845–3847;<br />

(c) D.S. McGuinness, K.J. Cavell, Organometallics 19 (2000) 741–<br />

748;<br />

(d) D.S. McGuinness, M.J. Green, K.J. Cavell, B.W. Skelton, A.H.<br />

White, J. Organomet. Chem. 565 (1998) 165–178.<br />

[9] (a) A. Fürstner, O.R. Thiel, C.W. Lehmann, Organometallics 21<br />

(2002) 331–335;<br />

(b) A. Fürstner, L. Ackermann, K. Beck, H. Hori, D. Koch, K.<br />

Langemann, M. Liebl, C. Six, W. Leitner, J. Am. Chem. Soc. 123<br />

(2001) 9000–9006;<br />

(c) A. Fürstner, O.R. Thiel, L. Ackermann, H.-J. Schanz, S.P. Nolan,<br />

J. Org. Chem. 65 (2000);<br />

(d) A. Fürstner, O.R. Thiel, N. Kindler, B. Bartkowska, J. Org.<br />

Chem. 65 (2000) 7990–7995.<br />

[10] (a) G. Altenhoff, R. Goddard, C.W. Lehmann, F. Glorius, J. Am.<br />

Chem. Soc. 126 (2004) 15195–15201;<br />

(b) G. Altenhoff, R. Goddard, C.W. Lehmann, F. Glorius, Angew.<br />

Chem., Int. Ed. 42 (2003) 3690–3693.<br />

[11] (a) T.W. Funk, J.M. Berlin, R.H. Grubbs, J. Am. Chem. Soc. 128<br />

(2006) 1840–1846;<br />

(b) J.P. Morgan, R.H. Grubbs, Org. Lett. 2 (2000) 3153–3155;<br />

(c) C.W. Bielawski, R.H. Grubbs, Angew. Chem., Int. Ed. 39 (2000)<br />

2903–2906.<br />

[12] (a) N. Marion, O. Navarro, J. Mei, E.D. Stevens, N.M. Scott, S.P.<br />

Nolan, J. Am. Chem. Soc. 128 (2006) 4101–4111;<br />

(b) S. Díez-González, H. Kaur, F.K. Zinn, E.D. Stevens, S.P. Nolan,<br />

J. Org. Chem. 70 (2005) 4784–4796;<br />

(c) H.M. Lee, T. Jiang, E.D. Stevens, S.P. Nolan, Organometallics 20<br />

(2001) 1255–1258;<br />

(d) J. Huang, E.D. Stevens, S.P. Nolan, J.L. Petersen, J. Am. Chem.<br />

Soc. 121 (1999) 2674–2678;<br />

(e) J. Huang, S.P. Nolan, J. Am. Chem. Soc. 121 (1999) 9889–<br />

9890;<br />

(f) J. Huang, G. Grasa, S.P. Nolan, Org. Lett. 1 (1999) 1307–<br />

1309.<br />

[13] (a) M.J. Schultz, S.S. Hamilton, D.R. Jensen, M.S. Sigman, J. Org.<br />

Chem. 70 (2005) 3343–3352;<br />

(b) C.M. Crudden, D.P. Allen, Coordin. Chem. Rev. 248 (2004)<br />

2247–2273;<br />

(c) M.K. Denk, J.M. Rodezno, S. Gupta, A.J. Lough, J. Organomet.<br />

Chem. 617–618 (2001) 242–253;<br />

(d) M. Regitz, Angew. Chem. 108 (1996) 791–794.<br />

[14] (a) S. Harkal, R. Jackstell, F. Nierlich, D. Ortmann, M. Beller, Org.<br />

Lett. 7 (2005) 541–544;<br />

(b) A. Zapf, M. Beller, Chem. Commun. (2005) 431–440;<br />

(c) C.-F. Huo, T. Zeng, Y.-W. Li, M. Beller, H. Jiao, Organometallics<br />

24 (2005) 6037–6042;<br />

(d) R. Jackstell, S. Harkal, H. Jiao, A. Spannenberg, C. Borgmann,<br />

D. Röttger, F. Nierlich, M. Elliot, S. Niven, K. Cavell, O. Navarro,<br />

M.S. Viciu, S.P. Nolan, M. Beller, Chem. Eur. J. 10 (2004) 3891–<br />

3900;<br />

(e) A.C. Frisch, F. Rataboul, A. Zapf, M. Beller, J. Organomet.<br />

Chem. 687 (2003) 403–409;<br />

(f) R. Jackstell, M.G. Andreu, A. Frisch, K. Selvakumar, A. Zapf,<br />

H. Klein, A. Spannenberg, D. Röttger, O. Briel, R. Karch, M. Beller,<br />

Angew. Chem., Int. Ed. 41 (2002) 986–989.<br />

[15] (a) E. Mas-Marzá, M. Poyatos, M. Sanaú, E. Peris, Organometallics<br />

23 (2004) 323–325;<br />

(b) M. Poyatos, E. Mas-Marzá, J.A. Mata, M. Sanaú, E. Peris, Eur.<br />

J. Inorg. Chem. (2003) 1215–1221;<br />

(c) M. Albrecht, R.H. Crabtree, J. Mata, E. Peris, Chem. Commun.<br />

(2002) 32–33.<br />

[16] (a) F.E. Hahn, C. Holtgrewe, T. Pape, M. Martin, E. Sola, L.A.<br />

Oro, Organometallics 24 (2005) 2203–2209;<br />

(b) J.R. Miecznikowski, R.H. Crabtree, Organometallics 23 (2004)<br />

629–631;<br />

(c) M. Albrecht, J.R. Miecznikowski, A. Samuel, J.W. Faller, R.H.<br />

Crabtree, Organometallics 21 (2002) 3596–3604;<br />

(d) A.C. Hillier, H.M. Lee, E.D. Stevens, S.P. Nolan, Organometallics<br />

20 (2001) 4246–4252.<br />

[17] (a) S. Burling, M.K. Whittlesey, J.M.J. Williams, Adv. Synth. Catal.<br />

347 (2005) 591–594;<br />

(b) P.L. Chiu, H.M. Lee, Organometallics 24 (2005) 1692–1702;


S. Enthaler et al. / Journal of Organometallic Chemistry 691 (2006) 4652–4659 4659<br />

(c) M.G. Edwards, R.F.R. Jazzar, B.M. Paine, D.J. Shermer, M.K.<br />

Whittlesey, J.M.J. Williams, D.D. Edney, Chem. Commun. (2004)<br />

90–91;<br />

(d) M. Poyatos, J.A. Mata, E. Falomir, R.H. Crabtree, E. Peris,<br />

Organometallics 22 (2003) 1110–1114;<br />

(e) A.A. Danopoulos, S. Winston, W.B. Motherwell, Chem. Commun.<br />

(2002) 1376–1377.<br />

[18] S. Kuhl, R. Schneider, Y. Fort, Organometallics 22 (2003) 4184–<br />

4186.<br />

[19] W. Baratta, J. Schütz, E. Herdtweck, W.A. Herrmann, P. Rigo, J.<br />

Organomet. Chem. 690 (2005) 5570–5575.<br />

[20] (a) F. Rataboul, A. Zapf, R. Jackstell, S. Harkal, T. Riermeier, A.<br />

Monsees, U. Dingerdissen, M. Beller, Chem. Eur. J. 10 (2004) 2983–<br />

2990;<br />

(b) A.C. Frisch, A. Zapf, O. Briel, B. Kayser, N. Shaikh, M. Beller,<br />

J. Mol. Catal. 214 (2004) 231–239;<br />

(c) A.M. Seayad, K. Selvakumar, A. Moballigh, M. Beller, Tetrahedron<br />

Lett. 44 (2003) 1679–1683;<br />

(d) A.C. Frisch, N. Shaikh, A. Zapf, M. Beller, Angew. Chem., Int.<br />

Ed. 41 (2002) 4056–4059;<br />

(e) K. Selvakumar, A. Zapf, A. Spannenberg, M. Beller, Chem. Eur.<br />

J. 8 (2002) 3901–3906;<br />

(f) K. Selvakumar, A. Zapf, M. Beller, Org. Lett. 4 (2002) 3031–3033;<br />

(g) R. Jackstell, A. Frisch, M. Beller, D. Röttger, M. Malaun, B.<br />

Bildstein, J. Mol. Catal. 185 (2002) 105–112.<br />

[21] P. Wasserscheid, W. Keim, Angew. Chem., Int. Ed. 39 (2000) 3773–<br />

3789.<br />

[22] (a) O. Pàmies, J.E. Bäckvall, Chem. Eur. J. 7 (2001) 5052–5058;<br />

(b) A. Aranyos, G. Csjernyik, K.J. Szabó, J.E. Bäckvall, Chem.<br />

Commun. (1999) 351–352;<br />

(c) M.L.S. Almeida, M. Beller, G.Z. Wang, J.E. Bäckvall, Chem.<br />

Eur. J. 2 (1996) 1533–1536;<br />

(d) G.Z. Wang, J.E. Bäckvall, J. Chem. Soc. Chem. Commun. (1992)<br />

337–339;<br />

(e) G.Z. Wang, J.E. Bäckvall, J. Chem. Soc. Chem. Commun. (1992)<br />

980–982;<br />

(f) R.L. Chowdhury, J.E. Bäckvall, J. Chem. Soc. Chem. Commun.<br />

(1991) 1063–1064;<br />

(g) B.N. Chaudret, D.J. Cole-Hamilton, R.S. Nohr, G. Wilkinson, J.<br />

Chem. Soc. Dalton Trans. (1977) 1546–1557;<br />

(h) J. Chatt, B.L. Shaw, A.E. Field, J. Chem. Soc. (1964) 3466–3475;<br />

(i) J.S.M. Samec, J.-E. Bäckvall, P.G. Andersson, P. Brandt, Chem.<br />

Soc. Rev. 35 (2006) 237–248.<br />

[23] (a) D.D. Tanner, G.E. Diaz, A. Potter, J. Org. Chem. 50 (1985)<br />

2149–2154;<br />

(b) M. Degueil-Castaing, A. Rahm, J. Org. Chem. 51 (1986) 1672–<br />

1676;<br />

(c) T. Ohkuma, M. Koizumi, H. Doucet, T. Pham, M. Kozawa, K.<br />

Murata, E. Katayama, T. Yokozawa, T. Ikariya, R. Noyori, J. Am.<br />

Chem. Soc. 120 (1998) 13529–13530;<br />

(d) J. Wu, J.-X. Ji, R. Guo, C.-H. Yeung, A.S.C. Chan, Chem. Eur.<br />

J. 9 (2003) 2963–2968.<br />

[24] E.C. Ashby, J.N. Argyropoulos, J. Org. Chem. 51 (1986) 3593–<br />

3597.<br />

[25] (a) L. Dahlenburg, R. Götz, Eur. J. Inorg. Chem. (2004) 888–905;<br />

(b) P.W.C. Cross, G.J. Ellames, J.S. Gibson, J.M. Herbert, W.J.<br />

Kerr, A.H. McNeill, T.W. Mathers, Tetrahedron 59 (2003) 3349–<br />

3358.<br />

[26] (a) L.Y. Kuo, D.M. Finigan, N.N. Tadros, Organometallics 22<br />

(2003) 2422–2425;<br />

(b) R.L. Elsenbaumer, H.S. Mosher, J. Org. Chem. 44 (1979) 600–<br />

604.<br />

[27] D. Klomp, T. Maschmeyer, U. Hanefeld, J.A. Peters, Chem. Eur. J.<br />

10 (2004) 2088–2093.<br />

[28] (a) L. Jafarpour, E.D. Stevens, S.P. Nolan, J. Organomet. Chem. 606<br />

(2000) 49–54;<br />

(b) S.V. Dzyuba, R.A. Bartsch, J. Heterocylic Chem. 38 (2001) 265–<br />

268.


6. Transfer hydrogenation<br />

6.2. Transfer hydrogenation of prochiral ketones with ruthenium catalysts<br />

6.2.2. New Ruthenium Catalysts for Asymmetric Transfer Hydrogenation<br />

of prochiral Ketones<br />

Stephan Enthaler, Bernhard Hagemann, Santosh Bhor, Gopinathan Anilkumar, Man Kin Tse,<br />

Bianca Bitterlich, Kathrin Junge, Giulia Erre, and Matthias Beller*, Adv. Synth. Catal. 2007,<br />

349, 853–860.<br />

Contributions<br />

Ligands 3a–3e were synthesized by Dr. S. Bhor, Dr. G. Anilkumar, Dr. M. K. Tse and B.<br />

Bitterlich. The experiments stated in table 4 entries 2 (cat. A), 3 (cat. B) and 6 (cat. A and<br />

cat. B) were carried out by Dr. B. Hagemann.<br />

134


New Ruthenium Catalysts for Asymmetric Transfer<br />

Hydrogenation of Prochiral Ketones<br />

DOI: 10.1002/adsc.200600475<br />

Stephan Enthaler, a Bernhard Hagemann, a Santosh Bhor, a Gopinathan Anilkumar, a<br />

Man Kin Tse, a Bianca Bitterlich, a Kathrin Junge, a Giulia Erre, a<br />

and Matthias Beller a, *<br />

a Leibniz-Institut für Katalyse e.V. an der <strong>Universität</strong> <strong>Rostock</strong>, Albert-Einstein-Str. 29a, 18059 <strong>Rostock</strong>, Germany<br />

Fax: (+49)-381-1281-5000; e-mail: matthias.beller@catalysis.de<br />

Received: September 15, 2006; Revised: February 5, 2007<br />

Abstract: Tridentate N,N,N-pyridinebisimidazolines<br />

have been studied as new ligands for the enantioselective<br />

transfer hydrogenation of prochiral ketones.<br />

High yields and excellent enantioselectivity up to<br />

> 99% ee have been achieved with an in situ generated<br />

catalytic system containing dichlorotris(triphenylphosphine)ruthenium<br />

and 2,6-bis-([4R,5R]-4,5diphenyl-4,5-dihydro-1H-imidazol-2-yl)-pyridine<br />

(3a) in the presence of sodium isopropoxide.<br />

Keywords: asymmetric transfer hydrogenation; ketones;<br />

phosphines; ruthenium; tridentate nitrogen<br />

ligands<br />

Enantiomerically pure alcohols have a wide range of<br />

applications, for example, building blocks and synthons<br />

for pharmaceuticals, agrochemicals, polymers,<br />

syntheses of natural compounds, auxiliaries, ligands<br />

and key intermediates in organic syntheses. [1] Within<br />

the different molecular transformations to chiral alcohols,<br />

transition metal-catalyzed reactions offer efficient<br />

and versatile strategies, such as addition of organometallic<br />

compounds to aldehydes, hydrosilylation,<br />

and hydrogenation of prochiral ketones. [2] From<br />

an economic and environmental point of view the<br />

asymmetric hydrogenation, in particular the transfer<br />

hydrogenation, represents a powerful tool for their<br />

synthesis because of its high atom economy and<br />

safety advantages. [3] Here, Noyori s ruthenium-based<br />

catalysts comprising chiral tosylated diamines constitute<br />

state-of-the-art transfer hydrogenation systems.<br />

[3d,4] Based on this seminal work an increasing<br />

number of ruthenium catalysts with chiral bidentate<br />

N,N-ligands were developed in the last decade. [3] Significantly<br />

fewer systems are known in which transfer<br />

of chiral information is promoted by tridentate ligands.<br />

[5,6] Up to now only a limited number of auspicious<br />

tridentate nitrogen-containing N,N,N ligands<br />

COMMUNICATIONS<br />

were established in the field of transfer hydrogenation.<br />

For example (R)-phenyl-ambox (1) [5] and different<br />

pyridinebisoxazoline (pybox) ligands (2) [6] have<br />

been applied for the reduction of acetophenone<br />

(Scheme 1).<br />

Scheme 1. N,N,N-Tridentate ligands.<br />

Recently, we reported the synthesis of a new class<br />

of chiral tridentate amines. [7] The preparation and<br />

tunability of these pyridinebisimidazolines (3) (socalled<br />

pybim ligands, Scheme 1) are easier and more<br />

flexible compared to the popular pyboxes, making the<br />

former a suitable ligand tool box for various asymmetric<br />

transformations. To date there is no report on<br />

the performance of these ligands in hydrogenation reactions.<br />

The resemblance between pybim (3) and<br />

pybox (2) stimulated our research to study the potential<br />

of this class of ligands in the transfer hydrogenation<br />

of aromatic and aliphatic ketones.<br />

In exploratory experiments, isopropyl alcohol-based<br />

transfer hydrogenation of acetophenone was examined<br />

using a simple in situ catalyst system composed<br />

of [RuCl 2ACHTUNGTRENUNG(C 6H 6)] 2, 2,6-bis-([4R,5R]-4,5-diphenyl-4,5-<br />

Adv. Synth. Catal. 2007, 349, 853 – 860 2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 853


COMMUNICATIONS<br />

dihydro-1H-imidazol-2-yl)-pyridine (3) and triphenylphosphine.<br />

To ensure complete formation of the<br />

active catalyst and avoid an induction period the reaction<br />

mixture was heated for 10 min at 100 8C in the<br />

presence of base followed by addition of acetophenone.<br />

First, studies for optimizing the reaction conditions<br />

were carried out with 1 mol% of pre-catalyst<br />

and 5 mol% of base. It is well-known that transfer hydrogenations<br />

are sensitive to the nature of the base.<br />

Thus, the influence of different bases on selectivity<br />

and conversion was investigated initially (Table 1).<br />

Best results were obtained for sodium isopropoxide<br />

and K 2CO 3 with conversions up to 95% and enantiomeric<br />

excesses up to 94% (Table 1, entries 1 and 8).<br />

Interestingly, NaOH and KOH the most commonly<br />

used bases for transfer hydrogenations gave only<br />

moderate enantioselectivity (78 % and 80%, Table 1,<br />

entries 3 and 4). In addition, we tested some organic<br />

nitrogen-containing systems such as DBU, DABCO,<br />

NEt 3, NACHTUNGTRENUNG(i-Pr) 2Et and pyridine, but only with NACHTUNGTRENUNG(i-<br />

Pr) 2Et did we obtain significant amounts of product<br />

in a reasonable time (Table 1, entry 10). Next, the<br />

concentration of sodium isopropoxide was varied at<br />

different temperatures. As expected, the increase of<br />

the amount of base led to an acceleration of reaction<br />

rate; however, this was accompanied by an unacceptable<br />

decrease of enantioselectivity (Table 1, entries 13–<br />

15). In contrast, improved ee is obtained by reducing<br />

the amount of base to a ruthenium-to-base ratio of 1<br />

to 0.5 (Table 1, entries 16 and 17). Notably, in the absence<br />

of base the transfer of hydrogen did not occur.<br />

Based on these results we investigated the behavior<br />

of the metal precursor.<br />

Applying different ruthenium sources such as<br />

[RuCl 2ACHTUNGTRENUNG(PPh 3) 3], [RuHClACHTUNGTRENUNG(PPh 3) 3], [8] RuCl 3·x H 2O,<br />

Ru 3(CO) 12 and RuACHTUNGTRENUNG(cod)(methylallyl) 2, lower conversion<br />

and/or poor selectivity were achieved. Nevertheless,<br />

[RuCl 2ACHTUNGTRENUNG(PPh 3) 3] showed an enantiomeric excess of<br />

> 99 %, which is to our knowledge the highest enantioselectivity<br />

in this model reaction for a chiral tridentate<br />

ligand (Table 1, entry 17). However, a slight<br />

excess of pybim 3a, 3 equivs. with respect to 1 equiv.<br />

Table 1. Transfer hydrogenation of acetophenone in the presence of pybim ligand 3a and different bases. [a]<br />

Entry Base Base:Metal Temp. [8C] Conv. [%] [b]<br />

Stephan Enthaler et al.<br />

ee [%] [b]<br />

1 NaO-i-Pr 5 100 80 94 (S)<br />

2 KO-t-Bu 5 100 99 9 (S)<br />

3 NaOH 5 100 79 78 (S)<br />

4 KOH 5 100 85 80 (S)<br />

5 LiOH 5 100 72 80 (S)<br />

6 K 3PO 4 5 100 65 70 (S)<br />

7 K 2HPO 4 5 100 19 96 (S)<br />

8 K 2CO 3 5 100 95 93 (S)<br />

9 Cs 2CO 3 5 100 45 67 (S)<br />

10 NACHTUNGTRENUNG(i-Pr) 2Et 5 100 18 90 (S)<br />

11 NaO-i-Pr 5 60 93 84 (S) [c]<br />

12 NaO-i-Pr 5 80 91 83 (S) [d]<br />

13 NaO-i-Pr 5 90 91 88 (S)<br />

14 NaO-i-Pr 50 90 98 87 (S)<br />

15 NaO-i-Pr 250 90 99 44 (S)<br />

16 NaO-i-Pr 0.5 100 88 95 (S)<br />

17 NaO-i-Pr 0.5 100 96 >99 (S) [e]<br />

18 NaO-i-Pr 0.5 110 58 84 (S)<br />

[a] 6<br />

Reaction conditions: in situ catalyst A {1.9 ”10 mol [RuCl2ACHTUNGTRENUNG(C6H6)] 2, 3.8 ” 10 6 mol ligand 3a, 3.8 ” 10 6 mol PPh3}; addition<br />

of the corresponding base: entries 1–13: 1.9 ” 10 5 mol, entry 14: 1.9 ” 10 4 mol, entry 15: 9.5 ”10 4 mol and for entries<br />

16–18: 1.9 ” 10 6 mol in 2.0 mL isopropyl alcohol, 10 min at corresponding temperature then addition of 3.8 ” 10 4<br />

mol acetophenone, 1 h at corresponding temperature.<br />

[b]<br />

Conversion and ee were determined by chiral GC (50 m Lipodex E, 95–2008C) analysis with diglyme as internal standard.<br />

[c]<br />

Reaction time 14 h.<br />

[d]<br />

Reaction time 4 h.<br />

[e] 6<br />

In situ catalyst B {3.8 ”10 mol [RuCl2ACHTUNGTRENUNG(PPh3) 3], 1.14 ” 10 5 mol 3a}. Conversion was determined by GC (30 m HP Agilent<br />

Technologies 50–300 8C) with diglyme as internal standard and ee was determined by chiral HPLC (Chiralcel OB-H,<br />

eluent: n-hexane/ethanol, 99:1, flow rate: 2 mL min 1 ) analysis.<br />

854 www.asc.wiley-vch.de 2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Synth. Catal. 2007, 349, 853 – 860


Ruthenium Catalysts for Asymmetric Transfer Hydrogenation<br />

of ruthenium, was necessary to achieve high enantioselectivity.<br />

To compare the difference of catalyst<br />

system A and catalyst system B, we investigated the<br />

conversion-time and the enantioselectivity-time dependency<br />

(Scheme 2). No significant change of enan-<br />

Scheme 2. Conversion-time and enantioselectivity-time behavior<br />

of catalyst A and catalyst B. Reaction conditions: in<br />

situ catalyst A (1.9 ” 10 6 mol [RuCl 2ACHTUNGTRENUNG(C 6H 6)] 2, 3.8 ” 10 6 mol<br />

ligand 3a, 3.8 ” 10 6 mol PPh 3)orin situ catalyst B (3.8 ”<br />

10 6 mol [RuCl 2ACHTUNGTRENUNG(PPh 3) 3], 1.14 ” 10 5 mol 3a); addition of<br />

sodium isopropoxide (1.9”10 6 mol) in 2.0 mL 2-propanol,<br />

10 min at 100 8C then addition of 3.8 ”10 4 mol acetophenone,<br />

reaction temperature: 100 8C. Conversion was determined<br />

by GC (30 m HP Agilent Technologies 50–300 8C)<br />

with diglyme as internal standard and ee was determined by<br />

chiral HPLC (Chiralcel OB-H, eluent: n-hexane/ethanol,<br />

99:1, flow rate: 2 mL min 1 ) analysis.<br />

tioselectivity was observed in the shown time frame,<br />

while after 24 h <strong>full</strong> racemization occurred with catalyst<br />

A. Noteworthy, the analysis of the conversiontime<br />

behavior proved a higher catalyst activity for catalyst<br />

B, while for catalyst A a slight deceleration was<br />

monitored. We assume a better stabilization of the<br />

active species by an excess of 3a and PPh 3.<br />

As shown in Table 2 we also examined different<br />

pybox ligands (2a–c), but only unsatisfying results<br />

were obtained demonstrating the advantage of 3a<br />

(Table 2, entries 1–3).<br />

Noyori et al. proposed a metal-ligand bifunctional<br />

mechanism for the hydride transfer process (“NH effect”).<br />

[4,13b] Hence, the NH group of the imidazoline<br />

rings could be involved in the selectivity transfer and<br />

increase the coordination affinity between substrate<br />

COMMUNICATIONS<br />

Table 2. Transfer hydrogenation of acetophenone in the<br />

presence of different pybox and pybim ligands. [a]<br />

Entry Ligand Time [h] Conv. [%] [b]<br />

ee [%] [b]<br />

1 2a 1 6 11 (R)<br />

2 2b 1 10 18 (S)<br />

3 2c 1 22 8 (R)<br />

4 3a 1 83 98 (S)<br />

5 3b 2 95 26 (R)<br />

6 3c 2 74 29 (S)<br />

7 3d 1 4 7 (R)<br />

8 3e 1 77 7 (S)<br />

[a] Reaction conditions: 3.8 ” 10 6 mol [RuCl2ACHTUNGTRENUNG(PPh 3) 3], 1.14 ”<br />

10 5 mol ligand, 3.8 ” 10 6 mol PPh 3, 1.9 ”10 6 mol NaOi-Pr,<br />

2.0 mL isopropyl alcohol, 10 min at 100 8C then addition<br />

of 3.8 ” 10 4 mol acetophenone.<br />

[b] Conversion and ee were determined by chiral GC (50 m<br />

Lipodex E, 95–200 8C) analysis with diglyme as internal<br />

standard.<br />

and catalyst. Due to the formation of a second binding<br />

site, a more favorable position for transferring the<br />

chiral information is attained. [5,9,13]<br />

In order to understand the role of the free NH<br />

functionality for our pybim ligand 3a, different substitutions<br />

at the imidazoline units were carried out. The<br />

corresponding monoprotected ligands 3b, c were synthesized<br />

by reaction of 3a with one equivalent of 3,5di-tert-butylbenzoic<br />

acid chloride or Ph 2P(O)Cl. We<br />

presumed that the resulting unsymmetrical complexes<br />

would direct the substrate to occupy a specific orientation<br />

in the transition state and thereby induce increased<br />

selectivity during catalysis. However, 3,5-(t-<br />

Bu) 2C 6H 3)CO (3b) or Ph 2PO substitution (3c) decreased<br />

the enantioselectivity in the model reaction<br />

significantly (Table 2, entries 5 and 6). To confirm the<br />

results of the monosubstituted ligands an exchange of<br />

both hydrogens with Boc (3d) orBz(3e) protecting<br />

groups was carried out. A further decrease of enantioselectivity<br />

was observed (Table 2, entries 7 and 8).<br />

Thus, there is a crucial necessity of both NH functionalities<br />

for obtaining high enantioselectivity in the<br />

transfer hydrogenation of acetophenone. Interestingly,<br />

the NH group is not necessary to achieve significant<br />

conversions (Table 2, entry 8), which is in contrast to<br />

previous reports by Noyori et al., for example catalysts<br />

containing N-dimethylamino alcohols are completely<br />

inactive compared to their N-monomethyl<br />

counterparts. [3d,13b] To estimate the influence of the ligands<br />

in more detail we varied also the phosphorus<br />

ligand part (Table 3). Among the different achiral ligands<br />

best results were obtained with PPh 3,(p-MeO-<br />

C 6H 4) 3P and (p-Me-C 6H 4) 3P (Table 3, entries 1, 3 and<br />

Adv. Synth. Catal. 2007, 349, 853 – 860 2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.asc.wiley-vch.de 855


COMMUNICATIONS<br />

Table 3. Influence of different phosphorus ligands on the transfer<br />

hydrogenation of acetophenone. [a]<br />

Entry P Ligand Temp.<br />

[8C]<br />

Conv.<br />

[%] [b]<br />

ee<br />

[%] [b]<br />

1 Ph 3P 100 80 94 (S)<br />

2 without 100 61 7 (S)<br />

3 (p-MeO-C 6H 4) 3P 90 91 87 (S) [c]<br />

4 (o-Me-C 6H 4) 3P 100 20 rac<br />

5 (p-Me-C 6H 4) 3P 100 41 97 (S)<br />

6 ACHTUNGTRENUNG[3,4-ACHTUNGTRENUNG(CF 3) 2-C 6H 3] 3P 90 24 66 (S) [c]<br />

7 ACHTUNGTRENUNG(p-F-C 6H 4) 3P 100 53 76 (S)<br />

8 Cy 3P 90 2 15 (S) [c]<br />

9 t-Bu 3P 100 1 12 (S)<br />

10 n-BuPAd 2 100 4 5 (S)<br />

11 ACHTUNGTRENUNG(i-PrO) 3P 100 28 75 (S)<br />

12 Ph 2PACHTUNGTRENUNG(CH 2)PPh 2 100 37 rac<br />

13 Ph 2PACHTUNGTRENUNG(CH 2) 2PPh 2 100 43 rac<br />

14 Ph 2PACHTUNGTRENUNG(CH 2) 5PPh 2 100 43 6 (S)<br />

15 Ph 2PACHTUNGTRENUNG(CH 2) 6PPh 2 100 15 37 (S)<br />

16 90 90 95 (S) [c]<br />

17 90 8 24 (S)<br />

[a] Reaction conditions: in situ catalyst {1.9”10 6 mol [RuCl2<br />

ACHTUNGTRENUNG(C 6H 6)] 2, 3.8 ” 10 6 mol 3a, 3.8 ”10 6 mol of the corresponding<br />

P ligand}, 1.9 ” 10 5 mol NaO-i-Pr, 2.0 mL isopropyl alcohol,<br />

10 min at described temperature then addition of 3.8 ”<br />

10 4 mol acetophenone, 1 h at described temperature.<br />

[b] Conversion and ee were determined by chiral GC (50 m Lipodex<br />

E, 95–200 8C) analysis with diglyme as internal standard.<br />

[c] Experiments also performed at 100 8C, but only low enantioselectivities<br />

or conversions were obtained.<br />

5). Methyl substitution in the o-position of arylphosphines<br />

decreased the enantioselectivity dramatically<br />

compared to that at the p-position (from 97 % ee to<br />

rac, entries 4 and 5). Moderate conversion and selectivity<br />

were obtained for electron-poor substituted arylphosphines<br />

(Table 3, entries 6 and 7). Interestingly,<br />

also more basic and sterically hindered alkylphosphines<br />

such as PCy 3,PACHTUNGTRENUNG(t-Bu) 3, and n-BuPAd 2 showed<br />

only low activity (Table 3, entries 8–10).<br />

Furthermore, we applied chelating arylphosphine ligands.<br />

For bis(diphenylphosphino)methane (DPPM)<br />

only disappointing conversion and selectivity were ob-<br />

Stephan Enthaler et al.<br />

tained (Table 3, entry 12). By increasing the number<br />

of CH 2 groups in the bridge an increase of the enantiomeric<br />

excess was detected (Table 3, entries 13–15),<br />

probably due to a weaker coordination of the second<br />

phosphine group to the ruthenium. In addition, to improve<br />

the enantioselectivity and to increase the enantiomeric<br />

differentiation in the shape of the catalysts,<br />

we investigated the influence of chiral phosphines.<br />

Therefore, we tested chiral monodentate ligands (S)-4<br />

and (S)-5 in the transfer hydrogenation of acetophenone<br />

in combination with pybim ligand 3a. Recently,<br />

we have demonstrated the successful application of<br />

such chiral monodentate 4,5-dihydro-3H-dinaphtho-<br />

[2,1-c;1’,2’-e]phosphepines in various asymmetric hydrogenations<br />

using molecular hydrogen. [10] Furthermore,<br />

Gladiali and co-workers reported previously<br />

the application of this ligand class in the rhodium-catalyzed<br />

asymmetric hydrogenation of C=C bonds. [21]<br />

However, only ligand (S)-4 showed comparable enantioselectivity<br />

to PPh 3 of 95% ee, while (S)-5 gave<br />

poor selectivity and conversion, due to a similar basicity<br />

to achiral alkylphosphines (Table 3, entry 16 and<br />

17).<br />

Next, we explored the influence of the concentration<br />

of PPh 3. The results indicated a necessity of<br />

1 equiv. of PPh 3 relating to 1 equiv. of ruthenium,<br />

while the reaction with more than 1 equiv. of PPh 3 or<br />

in the absence of PPh 3 resulted in a decrease of enantioselectivity<br />

(Table 3, entries 1 and 2). Noteworthy,<br />

the use of an excess of PPh 3 has no disordered effect<br />

on selectivity while a negative influence was reported<br />

for other catalytic systems when PPh 3 was not removed.<br />

[5,6] In analogy to Gimeno et al. [6,11] and Yu<br />

et al. [12] we assume a cis-coordination of the phosphine<br />

with respect to the N-N-N plane, which forms<br />

after removal of the cis-coordinated chlorides (with<br />

respect to each other) a highly selective vacancy for<br />

substrate coordination and chirality transfer.<br />

To demonstrate the usefulness of the novel catalysts<br />

we employed system A ([RuCl 2ACHTUNGTRENUNG(C 6H 6)] 2/pybim/PPh 3)<br />

and B ([RuCl 2ACHTUNGTRENUNG(PPh 3) 3]/pybim) in the asymmetric<br />

transfer hydrogenation of six aromatic and one aliphatic<br />

ketones (Table 4). In general, catalyst system<br />

A gave some higher enantioselectivities compared to<br />

catalyst B. Substituted acetophenones and propiophenone<br />

gave enantioselectivities up to 98% ee (Table 4,<br />

entries 1–5). In the case of methoxy-substitution the<br />

position of the substituent plays an important role,<br />

because ortho-substitution was favored (Table 4, entries<br />

3 and 4). A chloro substituent in the a-position<br />

to the carbonyl group proved to be problematic and<br />

deactivated both catalysts (Table 4, entry 6). Compared<br />

to aromatic ketones, aliphatic ketones are more<br />

challenging substrates. Nevertheless, 1-cyclohexylethanone<br />

was reduced by catalyst A in good yield and<br />

enantioselectivity (Table 4, entry 7).<br />

856 www.asc.wiley-vch.de 2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Synth. Catal. 2007, 349, 853 – 860


Ruthenium Catalysts for Asymmetric Transfer Hydrogenation<br />

Table 4. Transfer hydrogenation of prochiral ketones. [a]<br />

Entry Alcohol Catalyst A [b,c,d]<br />

Catalyst B [b,c,d]<br />

Conv. [%] ee [%] Conv. [%] ee [%]<br />

1 89 89 (S) > 99 85 (S)<br />

2 84 94 (S) > 99 89 (S)<br />

3 86 [e]<br />

Various methods ( 1 H, 13 C, 31 Pand 15 N NMR, COSY<br />

NMR, HMQC NMR and MS) were used for characterization<br />

of the pre-catalyst, but unfortunately no<br />

precise structure has been proven. The 31 P NMR spectrum<br />

of the pre-catalyst in CDCl 3 indicated a single<br />

compound, because a singlet appeared at 30 ppm<br />

(free PPh 3: 6 ppm and O=PPh 3: 27 ppm). Furthermore,<br />

the composition of the pre-catalyst was confirmed<br />

by HR-MS as [RuCl 2ACHTUNGTRENUNG(PPh 3)(3a)] (calculated<br />

mass: 953.17549, detected mass: 953.17572). However,<br />

the coordination abilities of ligand 3a are so far unclear,<br />

because a rapid exchange of NH protons was<br />

detected, which causes a broad signal in the 1 HNMR<br />

72 (S) 68 [f]<br />

4 >99 98 ( ) 99 g]<br />

5 >99 97 (S) 98 [e]<br />

6 12 [g]<br />

70 (R) < 2 [f]<br />

7 91 82 (S) 96 [f]<br />

74 (S)<br />

70 ( )<br />

87 (S)<br />

12 (R)<br />

[a] Reaction conditions: in situ catalyst A {1.9 ”10 6 mol [RuCl2ACHTUNGTRENUNG(C 6H 6)] 2, 3.8 ” 10 6 mol 3a, 3.8 ” 10 6 mol PPh 3), 1.9 ” 10 5 mol<br />

NaO-i-Pr, 2.0 mL isopropyl alcohol, 10 min at 100 8C then addition of 3.8 ” 10 4 mol substrate, 1 h at 100 8C.<br />

[b] Conversion and ee were determined by chiral GC (entry 1: 25 m Lipodex E, 80–180 8C; entry 2: 25 m Lipodex E, 100 8C;<br />

entry 3: 50 m Lipodex E, 90–105 8C; entry 4: 50 m Lipodex E, 90–180 8C; entry 5: 25 m Lipodex E, 90–1808C; entry 6:<br />

50 m Lipodex E, 95–180 8C; entry 7: 25 m Lipodex E, 100 8C) analysis with diglyme as internal standard.<br />

[c] In situ catalyst B {3.8 ”10 6 mol [RuCl2ACHTUNGTRENUNG(PPh 3) 3], 1.14 ” 10 5 mol 3a}.<br />

[d] The absolute configurations were determined by comparing the sign of specific rotation with reported data.<br />

[e] 4h.<br />

[f] 8h.<br />

[g] 24 h.<br />

COMMUNICATIONS<br />

72 (S)<br />

spectrum for the four protons adjacent to the nitrogen<br />

atoms and furthermore one signal for the corresponding<br />

carbons in the 13 C NMR spectrum.<br />

Next, we focused our attention on a deeper comprehension<br />

of the reaction mechanism. For metal-catalyzed<br />

transfer hydrogenation two general mechanisms<br />

are accepted, designated as direct hydrogen<br />

transfer via formation of a six-membered cyclic transition<br />

state composed of metal, hydrogen donor and acceptor,<br />

and the hydridic route, which is subdivided<br />

into two pathways, the monohydride and dihydride<br />

mechanism (Scheme 3). More specifically, the formation<br />

of monohydride-metal complexes promotes an<br />

Adv. Synth. Catal. 2007, 349, 853 – 860 2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.asc.wiley-vch.de 857


COMMUNICATIONS<br />

exclusive hydride transfer from carbon (donor) to carbonyl<br />

carbon (acceptor) (Scheme 3), whereas a hydride<br />

transfer via dihydride-metal complexes leads to<br />

Scheme 3. Monohydride and dihydride mechanisms for<br />

transfer hydrogenations.<br />

no accurate prediction of hydride resting state, because<br />

the former hydride was transferred to the carbonyl<br />

carbon (acceptor) as well as to the carbonyl<br />

oxygen (acceptor) (Scheme 3). Indications for both<br />

pathways (hydridic route) have been established by<br />

various research groups, when following the hydride<br />

transfer catalyzed by metal complexes, e.g., Ru, Rh or<br />

Ir. [13]<br />

Reaction of cyclopropyl phenyl ketone (“radical<br />

clock”-substrate) with isopropyl alcohol in the presence<br />

of 1 mol % catalyst A gave exclusively the corresponding<br />

cyclopropylphenyl alcohol (> 99% by<br />

1 H NMR). Apparently there is no radical reduction<br />

induced by the transition metal or by sodium alkoxides.<br />

[14] The second assumption is also confirmed by<br />

performing the reduction of acetophenone in the<br />

presence of base and in the absence of the ruthenium<br />

catalyst. Here, no product at all was detected.<br />

Next, we followed the incorporation of hydrogen<br />

from the donor molecule (isopropyl alcohol) into the<br />

product by applying a deuterated donor. [15] The precatalyst<br />

(1 mol %) is generated by stirring a solution<br />

of isopropyl alcohol-d 8, 0.5 equivs. of [RuCl 2ACHTUNGTRENUNG(C 6H 6)] 2,<br />

ligand 3a and PPh 3 for 16 h at 65 8C. Then, sodium<br />

Scheme 4. Deuterium incorporation into acetophenone catalyzed by catalyst system A.<br />

Stephan Enthaler et al.<br />

isopropoxide-d 7 was added and the solution was stirred<br />

for 10 min at 100 8C. The reaction mixture was<br />

charged with acetophenone and after 1 h compound<br />

13 was observed as main product (> 99 %) by<br />

1 H NMR (Scheme 4). [16] The result proved an exclusive<br />

transfer of the deuterium into the carbonyl<br />

group, therefore a C H activation of the substrate/<br />

product under the described conditions did not occur.<br />

Furthermore, this result rules out an enol formation<br />

in the catalytic cycle. [17]<br />

To specify the position and the nature of the transferred<br />

hydride, the reaction was performed with isopropyl<br />

alcohol-d 1 (hydroxy group deuterated) as solvent/donor<br />

and sodium isopropoxide as base under<br />

identical reaction conditions. In the transfer hydrogenation<br />

of acetophenone we obtained a mixture of two<br />

deuterated 1-phenylethanols (Scheme 4, 14a and<br />

14b).The ratio between 14a and 14b (85:15) indicated<br />

a specific migration of the hydride, albeit some scrambling<br />

was detected. [18] This was probably caused by rearrangement<br />

of the hydride complex, starting from<br />

HN Ru D via N=Ru(HD) to DN Ru H and subsequent<br />

transfer process into acetophenone yielding<br />

14b. In conclusion the incorporation is in agreement<br />

with the monohydride mechanism, implying the formation<br />

of a metal hydride species in the catalytic<br />

cycle (Scheme 3). Furthermore, this indication is confirmed<br />

by the above-mentioned influence of the NH<br />

groups. In conclusion, the transfer of hydrogen, in the<br />

case of catalyst A, is subdivided into the hydride<br />

transfer by the metal and the proton transfer by the<br />

NH group in analogy to the metal-ligand bifunctional<br />

catalysis. [13r]<br />

We demonstrated for the first time the successful<br />

application of chiral tridentate pyridinebisimidazoline<br />

ligands in the asymmetric ruthenium-catalyzed transfer<br />

hydrogenation of aliphatic and aromatic ketones.<br />

Enantioselectivities up to > 99% ee were obtained<br />

under optimized reaction conditions. Comparison experiments<br />

of 3a with monoprotected pybims and<br />

pybox ligands displayed the crucial influence of the<br />

free NH functionality. Mechanistic experiments indicated<br />

the transfer of hydrogen from the donor molecule<br />

into the substrate via a metal-ligand bifunctional<br />

mechanism.<br />

858 www.asc.wiley-vch.de 2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Synth. Catal. 2007, 349, 853 – 860


Ruthenium Catalysts for Asymmetric Transfer Hydrogenation<br />

Experimental Section<br />

General Remarks<br />

All manipulations were performed under an argon atmosphere<br />

using standard Schlenk techniques. Isopropyl alcohol<br />

was used without further purification (purchased from<br />

Fluka, dried over molecular sieves). Sodium isopropoxide<br />

was prepared by reacting sodium with isopropyl alcohol<br />

under an argon atmosphere (stock solution). All ketones<br />

were dried over CaH 2, distilled in vacuum and stored under<br />

argon, except 4’-methoxyacetophenone and phenacyl chloride,<br />

which were used without further purification.<br />

BuP(Ad) 2 [19] and N-phenyl-2-(di-tert-butylphosphino)-pyrrole<br />

[20] were synthesized according to literature protocols.<br />

General Procedure for the Transfer Hydrogenation of<br />

Ketones<br />

In a 10-mL Schlenk tube, the in situ catalyst was prepared<br />

by stirring a solution of [RuCl 2ACHTUNGTRENUNG(C 6H 5)] 2 (1.9 ” 10 6 mmol),<br />

ligand 3a (3.8 ” 10 6 mmol) and PPh 3 (3.8 ” 10 6 mmol) in<br />

1.0 mL isopropyl alcohol for 16 h at 65 8C. To this mixture<br />

sodium isopropoxide (1.9 ”10 5 mmol in 0.5 mL isopropyl alcohol<br />

(stock solution)) was added and the solution stirred at<br />

100 8C for 10 min. After addition of the corresponding<br />

ketone (0.38 mmol in 0.5 mL isopropyl alcohol (stock solution))<br />

the reaction mixture was stirred for 1 h at 100 8C. The<br />

solution was cooled to room temperature and filtered over a<br />

plug of silica. The conversion and ee were measured by GC<br />

without further manipulations.<br />

Acknowledgements<br />

This work has been supported by the State of Mecklenburg-<br />

Western Pomerania and the “Bundesministerium für Bildung<br />

und Forschung (BMBF)”. Additional support came from the<br />

BMBF excellence center “CELISCA” and the “Graduiertenkolleg<br />

1213 - Neue Methoden für Nachhaltigkeit in Katalyse<br />

und Technik” and Leibniz price of the DFG. We thank Mrs.<br />

M. Heyken, Mrs. C. Voss, Dr. C. Fischer, Mrs. S. Buchholz<br />

(all Leibniz-Institut für Katalyse e.V. an der <strong>Universität</strong> <strong>Rostock</strong>)<br />

and J. Schumacher (<strong>Universität</strong> <strong>Rostock</strong>) for their excellent<br />

technical and analytical support.<br />

References<br />

[1] a) I. C. Lennon, J. A. Ramsden, Org. Process Res. Dev.<br />

2005, 9, 110 – 112; b) J. M. Hawkins, T. J. N. Watson,<br />

Angew. Chem. 2004, 116, 3286 – 3290; c) R. Noyori, T.<br />

Ohkuma, Angew. Chem. 2001, 113, 40–75; d) M.<br />

Miyagi, J. Takehara, S. Collet, K. Okano, Org. Process<br />

Res. Dev. 2000, 4, 346 – 348; e) E. N. Jacobsen, A.<br />

Pfaltz, H. Yamamoto, (Eds.), Comprehensive Asymmetric<br />

Catalysis, Springer, Berlin, 1999; f) R. Noyori,<br />

Asymmetric Catalysis in Organic Synthesis, Wiley, New<br />

York, 1994.<br />

[2] a) M. Beller, C. Bolm, (Eds.), Transition Metals for Organic<br />

Synthesis, Wiley-VCH, Weinheim, 2 nd edn, 2004;<br />

b) B. Cornils, W. A. Herrmann, Applied homogeneous<br />

COMMUNICATIONS<br />

Catalysis with Organometallic Compounds, Wiley-<br />

VCH, Weinheim, 1996.<br />

[3] a) S. Gladiali, E. Alberico, Chem. Soc. Rev. 2006, 35,<br />

226 – 236; b) S. Gladiali, G. Mestroni, in: Transition<br />

Metals for Organic Synthesis, (Eds.: M. Beller, C.<br />

Bolm), Wiley-VCH, Weinheim, 2 nd edn, 2004, pp 145 –<br />

166; c) H.-U. Blaser, C. Malan, B. Pugin, F. Spindler,<br />

M. Studer, Adv. Synth. Catal. 2003, 345, 103 – 151; d) R.<br />

Noyori, S. Hashiguchi, Acc. Chem. Res. 1997, 30, 97–<br />

102; e) G. Zassinovich, G. Mestroni, S. Gladiali, Chem.<br />

Rev. 1992, 51, 1051 – 1069.<br />

[4] a) T. Ikariya, S. Hashiguchi, K. Murata, R. Noyori,<br />

Org. Synth. 2005, 82, 10 – 17; b) K. Matsumura, S. Hashiguchi,<br />

T. Ikariya, R. Noyori, J. Am. Chem. Soc. 1997,<br />

119, 8738 – 8739; c) S. Hashiguchi, A. Fujii, K.-J. Haack,<br />

K. Matsumura, T. Ikariya, R. Noyori, Angew. Chem.<br />

Int. Ed. 1997, 36, 288 – 290; d) K.-J. Haack, S. Hashiguchi,<br />

A. Fujii, T. Ikariya, R. Noyori, Angew. Chem. Int.<br />

Ed. 1997, 36, 285 – 288; e) N. Uematsu, A. Fujii, S. Hashiguchi,<br />

T. Ikariya, R. Noyori, J. Am. Chem. Soc. 1996,<br />

118, 4916 – 4917; f) A. Fujii, S. Hashiguchi, N. Uematsu,<br />

T. Ikariya, R. Noyori, J. Am. Chem. Soc. 1996, 118,<br />

2521 – 2522; g) J.-X. Gao, T. Ikariya, R. Noyori, Organometallics<br />

1996, 15, 1087 – 1089; h) J. Takehara, S. Hashiguchi,<br />

A. Fujii, S. Inoue, T. Ikariya, R. Noyori,<br />

Chem. Commun. 1996, 233 – 234; i) S. Hashiguchi, A.<br />

Fujii, J. Takehara, T. Ikariya, R. Noyori, J. Am. Chem.<br />

Soc. 1995, 117, 7562 – 7563.<br />

[5] Y. Jiang, Q. Jiang, X. Zhang, J. Am. Chem. Soc. 1998,<br />

120, 3817 – 3818.<br />

[6] D. Cuervo, M. P. Gamasa, J. Gimeno, Chem. Eur. J.<br />

2004, 10, 425 – 432.<br />

[7] a) S. Bhor, G. Anilkumar, M. K. Tse, M. Klawonn, C.<br />

Dçbler, B. Bitterlich, A. Grotevendt, M. Beller, Org.<br />

Lett. 2005, 7, 3393 – 3396; b) G. Anilkumar, S. Bhor,<br />

M. K. Tse, M. Klawonn, B. Bitterlich, M. Beller, Tetrahedron:<br />

Asymmetry 2005, 16, 3536 – 3561.<br />

[8] No product was obtained when [RuHClACHTUNGTRENUNG(PPh 3) 3]/3a was<br />

performed without addition of base.<br />

[9] Y. Shvo, D. Czarkie, Y. Rahamim, J. Am. Chem. Soc.<br />

1986, 108, 7400 – 7402.<br />

[10] a) K. Junge, G. Oehme, A. Monsees, T. Riermeier, U.<br />

Dingerdissen, M. Beller, Tetrahedron Lett. 2002, 43,<br />

4977 – 4980; b) K. Junge, G. Oehme, A. Monsees, T.<br />

Riermeier, U. Dingerdissen, M. Beller, J. Organomet.<br />

Chem. 2003, 675, 91 – 96; c) K. Junge, B. Hagemann, S.<br />

Enthaler, A. Spannenberg, M. Michalik, G. Oehme, A.<br />

Monsees, T. Riermeier, M. Beller, Tetrahedron: Asymmetry<br />

2004, 15, 2621 – 2631; d) K. Junge, B. Hagemann,<br />

S. Enthaler, G. Oehme, M. Michalik, A. Monsees, T.<br />

Riermeier, U. Dingerdissen, M. Beller, Angew. Chem.<br />

Int. Ed. 2004, 43, 5066 – 5069; e) B. Hagemann, K.<br />

Junge, S. Enthaler, M. Michalik, T. Riermeier, A. Monsees,<br />

M. Beller, Adv. Synth. Catal. 2005, 347, 1978 –<br />

1986; f) S. Enthaler, B. Hagemann, K. Junge, G. Erre,<br />

M. Beller, Eur. J. Org. Chem. 2006, 2912 – 2917.<br />

[11] V. Cadierno, M. P. Gamasa, J. Gimeno, L. Iglesias,<br />

Inorg. Chem. 1999, 38, 2874 – 2879.<br />

[12] H. Deng, Z. Yu, J. Dong, S. Wu, Organometallics 2005,<br />

24, 4110 – 4112.<br />

[13] For recent review see: a) J. S. M. Samec, J.-E. Bäckvall,<br />

P. G. Andersson, P. Brandt, Chem. Soc. Rev. 2006, 35,<br />

Adv. Synth. Catal. 2007, 349, 853 – 860 2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.asc.wiley-vch.de 859


COMMUNICATIONS<br />

237 – 248; b) T. Ikariya, K. Murata, R. Noyori, Org.<br />

Biomol. Chem. 2006, 4, 393 – 406, and cited references<br />

therein; c) H. Guan, M. Iimura, M. P. Magee, J. R.<br />

Norton, G. Zhu, J. Am. Chem. Soc. 2005, 127, 7805 –<br />

7814; d) M. Gómez, S. Janset, G. Muller, G. Aullón,<br />

M. A. Maestro, Eur. J. Inorg. Chem. 2005, 4341 – 4351;<br />

e) K. MuÇiz, Angew. Chem. 2005, 117, 6780 – 6785;<br />

Angew. Chem. Int. Ed. 2005, 44, 6622 – 6627; f) C. Hedberg,<br />

K. Källstrçm, P. I. Arvidsson, P. Brandt, P. G. Andersson,<br />

J. Am. Chem. Soc. 2005, 127, 15083 – 15090;<br />

g) A. S. Y. Yim, M. Wills, Tetrahedron 2005, 61, 7994 –<br />

8004; h) S. E. Clapham, A. Hadzovic, R. H. Morris,<br />

Coord. Chem. Rev. 2004, 248, 2201 – 2237; i) T. Koike,<br />

T. Ikariya, Adv. Synth. Catal. 2004, 346, 37–41; j) P.<br />

Brandt, P. Roth, P. G. Andersson, J. Org. Chem. 2004,<br />

69, 4885 – 4890; k) J.-W. Handgraaf, J. N. H. Reek, E. J.<br />

Meijer, Organometallics 2003, 22, 3150 – 3157; l) C. P.<br />

Casey, J. B. Johnson, J. Org. Chem. 2003, 68, 1998 –<br />

2001; m) C. A. Sandoval, T. Ohkuma, K. MuÇiz, R.<br />

Noyori, J. Am. Chem. Soc. 2003, 125, 13490 – 13503;<br />

n) R. Noyori, Angew. Chem. 2002, 114, 2108 – 2123;<br />

Angew. Chem. Int. Ed. 2002, 41, 2008 – 2022; o) C. P.<br />

Casey, S. W. Singer, D. R. Powell, R. K. Hayashi, M.<br />

Kavana, J. Am. Chem. Soc. 2001, 123, 1090 – 1100; p) O.<br />

Pàmies, J. E. Bäckvall, Chem. Eur. J. 2001, 7, 5052 –<br />

5058; q) C. S. Yi, Z. He, Organometallics 2001, 20,<br />

3641 – 3643; r) R. Noyori, M. Yamakawa, S. Hashiguchi,<br />

J. Org. Chem. 2001, 66, 7931 – 7944; s) M. Yamakawa,<br />

H. Ito, R. Noyori, J. Am. Chem. Soc. 2000, 122, 1466 –<br />

1478; t) D. A. Alonso, P. Brandt, S. J. M. Nordin, P. G.<br />

Andersson, J. Am. Chem. Soc. 1999, 121, 9580 – 9588;<br />

u) D. G. I. Petra, J. N. H. Reek, J.-W. Handgraaf, E. J.<br />

Meijer, P. Dierkes, P. C. J. Kamer, J. Brussee, H. E.<br />

Schoemaker, P. W. N. M. van Leeuwen, Chem. Eur. J.<br />

Stephan Enthaler et al.<br />

2000, 6, 2818 – 2829; v) A. Aranyos, G. Csjernyik, K. J.<br />

Szabó, J. E. Bäckvall, Chem. Commun. 1999, 351 – 352;<br />

w) M. L. S. Almeida, M. Beller, G. Z. Wang, J. E. Bäckvall,<br />

Chem. Eur. J. 1996, 2, 1533 – 1536.<br />

[14] a) D. D. Tanner, G. E. Diaz, A. Potter, J. Org. Chem.<br />

1985, 50, 2149 – 2154; b) M. Degueil-Castaing, A.<br />

Rahm, J. Org. Chem. 1986, 51, 1672 – 1676; c) T.<br />

Ohkuma, M. Koizumi, H. Doucet, T. Pham, M.<br />

Kozawa, K. Murata, E. Katayama, T. Yokozawa, T.<br />

Ikariya, R. Noyori, J. Am. Chem. Soc. 1998, 120,<br />

13529 – 13530; d) J. Wu, J.-X. Ji, R. Guo, C.-H. Yeung,<br />

A. S. C. Chan, Chem. Eur. J. 2003, 9, 2963 – 2968;<br />

e) E. C. Ashby, J. N. Argyropoulos, J. Org. Chem. 1986,<br />

51, 3593 – 3597.<br />

[15] a) L. Dahlenburg, R. Gçtz, Eur. J. Inorg. Chem. 2004,<br />

888 – 905; b) P. W. C. Cross, G. J. Ellames, J. M. Gibson,<br />

J. M. Herbert, W. J. Kerr, A. H. McNeill, T. W. Mathers,<br />

Tetrahedron 2003, 59, 3349 – 3358.<br />

[16] a) L. Y. Kuo, D. M. Finigan, N. N. Tadros, Organometallics<br />

2003, 22, 2422 – 2425; b) R. L. Elsenbaumer, H. S.<br />

Mosher, J. Org. Chem. 1979, 44, 600 – 604.<br />

[17] D. Klomp, T. Maschmeyer, U. Hanefeld, J. A. Peters,<br />

Chem. Eur. J. 2004, 10, 2088 – 2093.<br />

[18] The ratio between 14a and 14b was determined by<br />

1<br />

H NMR based on the integrals of the signals for the<br />

CH3 groups.<br />

[19] A. Ehrentraut, A. Zapf, M. Beller, Synlett 2000, 1589 –<br />

1592.<br />

[20] A. Zapf, R. Jackstell, F. Rataboul, T. Riermeier, A.<br />

Monsees, C. Fuhrmann, N. Shaikh, U. Dingerdissen, M.<br />

Beller, Chem. Commun. 2004, 38 – 39.<br />

[21] E. Alberico, I. Nieddu, R. Taras, S. Gladiali, Helv.<br />

Chim. Acta 2006, 89, 1716 – 1729.<br />

860 www.asc.wiley-vch.de 2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Synth. Catal. 2007, 349, 853 – 860


7. Summary 143<br />

7. Summary<br />

7.1. Application of monodentate phosphines in asymmetric hydrogenations<br />

Since previous works have proven an excellent behaviour of ligands containing 4,5–dihydro–<br />

3H–dinaphtho[2,1–c;1´,2´–e]phosphepine motif in asymmetric hydrogenation of α–amino<br />

acid precursors, itaconic acid derivatives and β–ketoesters, we extended the scope and<br />

limitation in the reduction of unsaturated C–C bonds. The rhodium–catalyzed asymmetric<br />

hydrogenation of different N–acyl enamides to give N–acyl protected optically active amines<br />

has been examined for the first time in detail in the presence of chiral monodentate 4,5–<br />

dihydro–3H–dinaphtho[2,1–c;1´,2´–e]phosphepine (Scheme 1). The enantioselectivity is<br />

largely dependent on the nature of the substituent at the phosphorous atom and the enamide<br />

structure. Under optimized conditions up to 95% ee and catalyst activity up to 2000 h -1 have<br />

been achieved.<br />

P R<br />

(L)<br />

HN<br />

O<br />

[Rh(cod) 2]BF 4 + 2.1 (L)<br />

H2 (2.5 bar),<br />

toluene, 30 °C, 6 h<br />

R1 R1 Scheme 1. Asymmetric hydrogenation of N–acyl enamides.<br />

R 2<br />

R 2<br />

HN<br />

O<br />

up to 95% ee<br />

Furthermore the potential of presented monophosphine ligands were shown in the rhodium–<br />

catalyzed asymmetric hydrogenation of various β–dehydroamino acid derivatives to give<br />

optically active β–amino acids (Scheme 2). Here enantioselectivities up to 94% ee were<br />

obtained after optimization of reaction conditions. Noteworthy, contradictory performance<br />

was observed for the E– and Z–isomer and different reaction conditions were necessary to<br />

achieve best enantioselectivity. In addition a switch of product configuration was observed<br />

depending on the nature of the double bond, which has been scarcely reported.<br />

P R<br />

(L)<br />

R 1<br />

O<br />

R 2<br />

NH<br />

O<br />

OR 3<br />

[Rh(cod) 2]BF 4 + 2.1 (L)<br />

H 2<br />

E-isomer (R)-product<br />

Z-isomer (S)-product<br />

Scheme 2. Asymmetric hydrogenation of β–dehydroamino acid derivates.<br />

R 1<br />

O<br />

R 2<br />

NH<br />

*<br />

O<br />

OR 3<br />

up to 94% ee


7. Summary 144<br />

Based on the pioneering work of Feringa, de Vries, Minnaard and co–workers the rhodium–<br />

catalyzed asymmetric hydrogenation of different acyclic and cyclic enol carbamates to yield<br />

optically active carbamates has been studied in the presence of chiral monodentate ligands<br />

based on a 4,5–dihydro–3H–dinaphtho[2,1–c;1´,2´–e]phosphepine scaffold (Scheme 3).<br />

Various reaction parameters have been investigated in detail and enantioselectivities up to<br />

96% ee have been achieved with an in situ catalyst. Noteworthy, high thermal stability of the<br />

catalyst in the range of 10–90 °C was noticed with respect to enantioselectivity (94–96% ee).<br />

P R<br />

(L)<br />

N<br />

O O [Rh(cod) 2]BF4 + 2.1 (L)<br />

MeOH, 24 h, H 2 (25 bar)<br />

Scheme 3. Asymmetric hydrogenation of enol carbamates.<br />

7.2. Transfer hydrogenation<br />

N<br />

O O<br />

up to 96% ee<br />

In situ generated homogeneous iron complexes catalyze the transfer hydrogenation of<br />

aliphatic and aromatic ketones utilizing 2–propanol as hydrogen donor in the presence of<br />

base. The influence of different reaction parameters on the catalytic activity was investigated<br />

in detail applying a three component catalyst system, composed of an iron salt, 2,2´:6´,2´´–<br />

terpyridine and PPh3 (Scheme 4). The scope and limitations of the described catalyst have<br />

been shown in the reduction of eleven different ketones. In most cases high conversion and<br />

excellent chemoselectivity are achieved. Mechanistic studies indicate a monohydride reaction<br />

pathway for the homogeneous iron catalyst.<br />

O<br />

1/3 Fe 3(CO) 12/terpy/PPh 3<br />

Na-2-OPr, 2-PrOH, 100 °C, 7 h<br />

Scheme 4. Homogeneous iron catalyst for reduction of ketones.<br />

Furthermore, for the first time in situ generated iron porphyrins have been applied as<br />

homogeneous catalysts for the transfer hydrogenation of ketones. Using 2–propanol as<br />

hydrogen source various ketones are reduced to the corresponding alcohols in good to<br />

OH


7. Summary 145<br />

excellent yield and selectivity. Under optimized reaction conditions high catalyst turnover<br />

frequencies up to 642 h -1 were achieved (Scheme 5).<br />

R<br />

N<br />

R<br />

NH N<br />

R<br />

HN<br />

R<br />

1<br />

O<br />

1/3 Fe 3(CO) 12/1<br />

Na-2-OPr, 2-PrOH, 100 °C, 7 h<br />

OH<br />

conversions up to >99%<br />

TOF up to 642 h -1<br />

Scheme 5. Transfer hydrogenation of ketones in the presence of biomimetic iron porphyrin<br />

catalysts.<br />

In situ generated iron porphyrins have been furthermore applied as homogeneous catalysts for<br />

the transfer hydrogenation of α–substituted ketones. Using 2–propanol as hydrogen donor<br />

various hydroxyl-protected 1,2–hydroxyketones are reduced to the corresponding mono<br />

protected 1,2–diols in good to excellent yield. Under optimized reaction conditions catalyst<br />

turnover frequencies up to 2500 h -1 have been achieved (Scheme 6).<br />

R<br />

NH N<br />

N<br />

R<br />

R<br />

HN<br />

R<br />

1<br />

R 1<br />

O<br />

R 2<br />

OR 3<br />

FeCl 2/1<br />

NaOH, 2-PrOH, 100 °C<br />

R 1<br />

OH<br />

R 2<br />

OR 3<br />

conversions up to >99%<br />

TOF up to 2500 h -1<br />

Scheme 6. Transfer hydrogenation of α–substituted ketones in the presence of biomimetic<br />

iron porphyrin catalysts.<br />

Novel ruthenium carbene complexes have been in situ generated and tested for the transfer<br />

hydrogenation of ketones. Applying Ru(cod)(methylallyl)2 in the presence of imidazolium<br />

salts in 2–propanol and sodium–2–propanolate as base, turnover frequencies up to 346 h -1<br />

have been obtained for the reduction of acetophenone (Scheme 7). A comparative study<br />

involving ruthenium carbene and ruthenium phosphine complexes demonstrated higher<br />

activity of ruthenium carbene complexes.


L =<br />

Cl -<br />

N +<br />

N<br />

7. Summary 146<br />

O OH<br />

[Ru(cod)methylallyl2]/L Na-2-OPr, 2-PrOH, 100 °C<br />

conversion up to >99%<br />

TOF up to 346 h -1<br />

Scheme 7. Transfer hydrogenation of ketones in the presence of ruthenium carbene catalysts.<br />

Tridentate N,N,N–pyridinebisimidazolines have been studied as new ligands for the<br />

enantioselective transfer hydrogenation of prochiral ketones. High yield and excellent<br />

enantioselectivity up to >99% ee have been achieved with an in situ generated catalytic<br />

system containing dichlorotris(triphenylphosphine)ruthenium and 2,6–bis–([4R,5R]–4,5–<br />

diphenyl–4,5–dihydro–1H–imidazol–2–yl)–pyridine in the presence of sodium 2–propanolate<br />

(Scheme 8).<br />

L =<br />

H H<br />

Ph<br />

N<br />

N<br />

N<br />

Ph<br />

N N<br />

Ph<br />

Ph<br />

O<br />

[RuCl 2(PPh 3) 3]/3L<br />

Na-2-OPr, 2-PrOH, 100 °C, 1 h<br />

OH<br />

*<br />

up to >99% ee<br />

Scheme 8. New ruthenium catalysts for asymmetric transfer hydrogenation of prochiral<br />

ketones.


Abstract<br />

Abstract<br />

Die vorliegende Dissertation beschäftigte sich mit der Anwendung von monodentaten<br />

Phosphinliganden in der asymmetrischen Hydrierung von C–C Doppelbindungen. Dabei<br />

konnte die Wirksamkeit von Rhodiumkatalysatoren, die Phosphinliganden mit der 4,5–<br />

Dihydro–3H–dinaphtho[2,1–c;1´,2´–e]phosphepin–Struktureinheit enthielten, gezeigt werden.<br />

Verschiedenste prochirale Substratklassen, wurden mit exzellenten Enantioselektivitäten von<br />

bis zu 96% ee zu den entsprechenden ungesättigten Derivaten reduziert. Ein zweiter<br />

Schwerpunkt der Arbeit lag in der Entwicklung neuartiger eisenbasierter Hydrierkatalysatoren<br />

und deren Anwendung in der Transferhydrierung von Ketonen, wobei Katalysatoraktivitäten<br />

von bis zu 2500 h -1 (TOF) erzielt werden konnten.<br />

This dissertation deals with the application of monodentate phosphine ligands in asymmetric<br />

hydrogenation of C–C double bonds. The usefulness of rhodium catalysts containing<br />

phosphines based on the 4,5–Dihydro–3H–dinaphtho[2,1–c;1´,2´–e]phosphepine scaffold was<br />

demonstrated in the reduction of various prochiral substrates, since excellent<br />

enantioselectivities up to 96% ee were achieved. Furthermore, the properties of iron–based<br />

catalysts were studied in the field of transfer hydrogenation. Here high catalyst activities up to<br />

2500 h -1 were obtained in the reduction of ketones.


PERSONAL INFORMATION<br />

Surname: Enthaler<br />

First name: Stephan<br />

Curriculum Vitae<br />

Curriculum Vitae<br />

Current adress: Thomas-Müntzer-Platz 21<br />

Nationality: German<br />

Birthday: 27.02.1980<br />

18057 <strong>Rostock</strong> (Germany)<br />

Birthplace: Hagenow (Germany)<br />

EDUCATION<br />

1986 – 1990 Elementary school “Ludwig-Reinhard-Schule“ in Boizenburg/Elbe<br />

1990 – 1992 August-Bebel-Schule in Boizenburg/Elbe<br />

1992 – 1998 Gymnasium Boizenburg<br />

1998 Abitur (degree: General qualification for university entrance)<br />

1999 – 2004 Course of studies: Chemistry at the University of <strong>Rostock</strong> (Diplom-Chemie)<br />

2003 – 2004 Master thesis (Diplomarbeit): “Synthesis and application of novel monodentate<br />

phosphine ligands in asymmetric hydrogenations“ under the supervision of<br />

Prof. M. Beller<br />

2004 Degree: Diplom-Chemiker (Master of Science)<br />

since 2004 PhD thesis at the University of <strong>Rostock</strong> under the supervision of Prof. M.<br />

Beller<br />

RESEARCH EXPERIENCE<br />

2001 Student assistant in the group of Prof. R. Mietchen at the University of<br />

<strong>Rostock</strong>: “Phase transfer catalysis assisted by ultrasound“


TEACHING EXPERIENCE<br />

Curriculum Vitae<br />

2002 Student assistant at the University of <strong>Rostock</strong>: “Practical course for medical<br />

students“<br />

RELATED PROFESSIONAL EXPERIENCE<br />

2002 Student assistant at the University of <strong>Rostock</strong>: “IUPAC Conference on<br />

Chemical Thermodynamics (ICCT-2002)”<br />

DEPARTMENTAL ACTIVITIES<br />

2000 – 2003 Representative of the chemistry student council at the University of <strong>Rostock</strong><br />

PROFESSIONAL ASSOCIATIONS<br />

GDCh (Gesellschaft Deutscher Chemiker)<br />

MILITARY SERVICE<br />

1998 – 1999 Basic military service: 3./183 and 1./183 armoured battalion in Boostedt


PUBLICATIONS<br />

JOURNAL CONTRIBUTIONS<br />

1. “Synthesis of chiral monodentate binaphthophosphepine ligands and their application<br />

in asymmetric hydrogenation“ K. Junge, B. Hagemann, S. Enthaler, A. Spannenberg,<br />

M. Michalik, G. Oehme, A. Monsees, T. Riermeier, M. Beller, Tetrahedron:<br />

Asymmetry 2004, 15, 2621–2631.<br />

2. “Enantioselective hydrogenation of β–ketoesters with monodentate ligands” K. Junge,<br />

B. Hagemann, S. Enthaler, G. Oehme, M. Michalik, A. Monsees, T. Riermeier, U.<br />

Dingerdissen, M. Beller, Angew. Chem. Int. Ed. 2004, 43, 6150; Angew. Chem. 2004,<br />

116, 5176–5179; Angew. Chem. Int. Ed. 2004, 43, 5066–5069.<br />

3. “A general method fort the enantioselective hydrogenation of β–ketoesters using<br />

monodentate binaphthophosphepine ligands“ B. Hagemann, K. Junge, S. Enthaler, M.<br />

Michalik, T. Riermeier, A. Monsees, M. Beller, Adv. Synth. Catal. 2005, 1978–1986.<br />

4. “Enantioselective rhodium–catalyzed hydrogenation of enamides in the presence of<br />

chiral monodentate phosphanes” S. Enthaler, B. Hagemann, G. Erre, K. Junge, M.<br />

Beller, Eur. J. Org. Chem. 2006, 2912–2917.<br />

5. “An environmentally benign process for the hydrogenation of ketones with<br />

homogeneous iron catalysts” S. Enthaler, B. Hagemann, G. Erre, K. Junge, M. Beller<br />

Chem. Asian. J. 2006, 1, 598–604.<br />

6. “Biomimetic transfer hydrogenation of ketones with iron porphyrin catalysts” S.<br />

Enthaler, G. Erre, M. K. Tse, K. Junge, M. Beller, Tetrahedron Lett. 2006, 47, 8095–<br />

8099.<br />

7. “Efficient transfer hydrogenation of ketones in the presence of ruthenium Nheterocyclic<br />

carbene catalysts” S. Enthaler, R. Jackstell, B. Hagemann, K. Junge, G.<br />

Erre, M. Beller, J. Organomet. Chem. 2006, 691, 4652–4659.<br />

8. “Synthesis and catalytic application of novel binaphthyl–derived phosphorous<br />

ligands” K. Junge, B. Hagemann, S. Enthaler, G. Erre, M. Beller, ARKIVOC, 2007, 5,<br />

50–66.<br />

9. “Development of practical rhodium phosphine catalysts for the hydrogenation of β–<br />

dehydroamino acid derivates” S. Enthaler, G. Erre, K. Junge, J. Holz, A. Börner, E.<br />

Alberico, I. Nieddu, S. Gladiali, M. Beller, Org. Process Res. Dev. 2007, 11, 568–577.<br />

10. “New ruthenium catalysts for asymmetric transfer hydrogenation of ketones” S.<br />

Enthaler, B. Hagemann, S. Bhor, G. Anilkumar, M. K. Tse, B. Bitterlich, K. Junge, G.<br />

Erre, M. Beller Adv. Synth. Catal. 2007, 349, 853–860.


11. “Enantioselective rhodium-catalyzed hydrogenation of enol carbamates in the<br />

presence of monodentate phosphines” S. Enthaler, G. Erre, K. Junge, D. Michalik, A.<br />

Spannenberg, S. Gladiali, M. Beller, Tetrahedron: Asymmetry, 2007, 18, 1288–1298.<br />

12. “Novel rhodium catalyst for asymmetric hydroformylation of styrene: Study of<br />

electronic and steric effects of phosphorus seven–membered ring ligands” G. Erre, S.<br />

Enthaler, K. Junge, S. Gladiali, M. Beller, J. Mol. Catal. A, 2008, 280, 148–155.<br />

13. “Synthesis and application of chiral monodentate phosphines in asymmetric<br />

hydrogenation” G. Erre, S. Enthaler, K. Junge, S. Gladiali, M. Beller, Coord. Chem.<br />

Rev. 2008, in print.<br />

14. “2–Hydroxy– and 2–amino–functional arylphosphines – syntheses, reactivity and use<br />

in catalysis” J. Heinicke, W. Wawrzyniak, N. Peulecke, B. R. Aluri, M. K.<br />

Kindermann, P. G. Jones, S. Enthaler, M. Beller, Phosphorus, Sulfur, Silicon and Rel.<br />

Elements, 2008, in print.<br />

15. “Synthesis of 1,2-diols via biomimetic transfer hydrogenation with iron porphyrin<br />

catalysts” S. Enthaler, B. Spilker, G. Erre, M. K. Tse, K. Junge, M. Beller,<br />

Tetrahedron, in print.<br />

16. “Iron–catalyzed Enantioselective Hydrosilylation of Ketones��”<br />

N. S. Shaikh, S.<br />

Enthaler, K. Junge, M. Beller, Angew. Chem. 2008, accepted.<br />

17. “Sustainable metal catalysis with iron – from rust to a rising star?” S. Enthaler, K.<br />

Junge, M. Beller, Angew. Chem. 2008, accepted.<br />

18. “Synthesis of new monodentate phosphoramidites and their application in iridiumcatalyzed<br />

asymmetric hydrogenations” G. Erre, K. Junge, S. Enthaler, D. Addis, D.<br />

Michalik, A. Spannenberg, M. Beller, Chem. Asian J. 2008, submitted.<br />

POSTER CONTRIBUTIONS<br />

1. “First enantioselective hydrogenation of β–ketoesters using monodentate ligands“ K.<br />

Junge, B. Hagemann, S. Enthaler, M. Beller, A. Monsees, T. Riermeier, ORCHEM,<br />

Bad Nauheim (Germany), 9–11 September 2004.<br />

2. “Enantioselective hydrogenation of β–ketoesters using monodentate ligands“ S.<br />

Enthaler, K. Junge, B. Hagemann, A. Monsees, T. Riermeier, M. Beller, COST-D24<br />

STEREOCAT2004, Venedig (Italy), 16–19 September 2004.<br />

3. “Enantioselective hydrogenation of β−ketoesters using monodentate phosphepine<br />

ligands“ B. Hagemann, K. Junge, S. Enthaler, A. Monsees, T. Riermeier, M. Beller, 8.<br />

SFB Symposium FB 380, FZ-Jülich (Germany), 7–8 Oktober 2004.<br />

4. “Application of monodentate binaphthophosphepine ligands in asymmetric<br />

hydrogenations“ S. Enthaler, K. Junge, B. Hagemann, A. Monsees, T. Riermeier, M.


Beller, XXXVIII. Jahrestreffen Deutscher Katalytiker, Weimar (Germany), 16–18<br />

March 2005.<br />

5. “Application of monodentate binaphthophosphepine ligands in asymmetric<br />

hydrogenations“ S. Enthaler, K. Junge, B. Hagemann, A. Monsees, T. Riermeier, M.<br />

Beller, Green Chemistry: Development of Sustainable Processes, <strong>Rostock</strong> (Germany),<br />

30 March – 1 April 2005.<br />

6. “First enantioselective hydrogenation of β–ketoesters using monodentate ligands” K.<br />

Junge, S. Enthaler, B. Hagemann, A. Monsees, T. Riermeier, M. Beller, Green<br />

Chemistry: Development of Sustainable Processes, <strong>Rostock</strong> (Germany), 30 March – 1<br />

April 2005.<br />

7. “Synthesis of monodentate binaphthophosphepine ligands and their application in<br />

asymmetric hydrogenation” B. Hagemann, K. Junge, S. Enthaler, G. Erre, A.<br />

Monsees, T. Riermeier, M. Beller, The 2005 European Younger Chemists´<br />

Conference, Brno (Czech Republic), 31 August – 3 September 2005.<br />

8. “Novel axially–chiral monodentate P–donor ligands for asymmetric catalysis“ G. Erre,<br />

K. Junge, S. Enthaler, B. Hagemann, F. Marras, S. Gladiali, M. Beller, XXXIX.<br />

Jahrestreffen Deutscher Katalytiker, Weimar (Germany), 15–17 March 2006.<br />

9. “Novel axially–chiral monodentate P–donor ligands for asymmetric catalysis“ G. Erre,<br />

K. Junge, S. Enthaler, B. Hagemann, F. Marras, S. Gladiali, M. Beller, XXII<br />

International Conference on Organometallic Chemistry (ICOMC 2006), Zaragoza<br />

(Spain), 23–28 July 2006.<br />

10. “Enantioselective rhodium–catalyzed hydrogenation of enamides in the presence of<br />

chiral monodentate phosphines“ S. Enthaler, B. Hagemann, K. Junge, G. Erre, M.<br />

Beller, Bayer–Doktorandenkurs, Leverkusen–Monheim–Wuppertal (Germany), 20–25<br />

August 2006.<br />

11. “Enantioselective hydrogenation of enamides and enol carbamates with monodentate<br />

binaphthophosphepines” S. Enthaler, G. Erre, K. Junge, M. Beller, 1 st European<br />

Chemistry Congress (1 st ECC), Budapest (Hungary), 26 August – 01 September 2006.<br />

12. “Synthesis of chiral binaphthophosphepine ligands and their application in asymmetric<br />

catalysis” G. Erre, K. Junge, B. Hagemann, S. Enthaler, M. Beller, 4 th International<br />

Forum Life Science Automation, <strong>Rostock</strong> (Germany), 14–15 September 2006.<br />

13. “Synthesis and catalytic applications of novel binaphthyl–derived phosphorous<br />

ligands” K. Junge, G. Erre, S. Enthaler, M. Beller, 40. Jahrestreffen Deutscher<br />

Katalytiker, Weimar (Germany), 14–16 March 2007.<br />

14. “Rhodium phosphine catalysts for the hydrogenation of β–dehydroamino acid<br />

derivates” S. Enthaler, G. Erre, K. Junge, J. Holz, A. Börner, E. Alberico, I. Nieddu, S.<br />

Gladiali, M. Beller, Heidelberg Forum of Molecular Catalysis, Heidelberg (Germany),<br />

22 June 2007.


15. “Asymmetric hydrogenation of enamides – from benchmark substrates to indol<br />

alkaloids” S. Enthaler, G. Erre, K. Junge, M. Beller, 15 th European Symposium on<br />

Organic Chemistry, Dublin (Ireland), 8–13 July, 2007.<br />

16. “Rhodium phosphine catalysts for the hydrogenation of β–dehydroamino acid<br />

derivates” S. Enthaler, G. Erre, K. Junge, J. Holz, A. Börner, E. Alberico, I. Nieddu, S.<br />

Gladiali, M. Beller, 10 th JCF–Frühjahrssymposium, <strong>Rostock</strong> (Germany), 27–29<br />

March 2008.<br />

17. “New phosphine–Ligands for palladium–catalyzed coupling reactions” T. Schulz, S.<br />

Enthaler, C. Torborg, T. Schareina, A. Zapf, M. Beller, 10 th JCF–<br />

Frühjahrssymposium, <strong>Rostock</strong> (Germany), 27–29 March 2008.<br />

PATENTS<br />

1. “Chirale Eisen–Phosphan–Katalysatoren und deren Verwendung in selektiven<br />

Hydrierungsreaktionen“ S. Enthaler, K. Junge, M. Beller, DE, submitted.<br />

2. „Chirale Eisen–Katalysatoren mit stickstoffhaltigen Liganden und deren Verwendung<br />

in selektiven Hydrierungsreaktionen“ S. Enthaler, K. Junge, M. Beller, DE, submitted.<br />

BOOK CONTRIBUTIONS<br />

“Applications of iron catalysts in reduction chemistry” S. Enthaler, K. Junge, M. Beller in<br />

B. Plietker (Ed.) “Iron catalysis in organic chemistry“ Wiley–VCH, Weinheim, 2008.<br />

TALKS<br />

1. “Development of new monodentate phosphine ligands for asymmetric hydrogenation”<br />

S. Enthaler, COST–D24 STEREOCAT2005, Barcelona (Spain), 15–18 September<br />

2005.<br />

2. “Development of new monodentate phosphine ligands for asymmetric hydrogenation”<br />

S. Enthaler, BASF–IfOK–Symposium, Ludwigshafen (Germany), 14–16 November<br />

2005.


Eidesstattliche Erklärung<br />

Ich versichere hiermit an Eides statt, dass ich die vorliegende Arbeit selbstständig angefertigt<br />

und ohne fremde Hilfe verfasst habe, keine außer den von mir angegebenen Hilfsmitteln und<br />

Quellen dazu verwendet habe und die den benutzten Werken inhaltlich und wörtlich<br />

entnommenen Stellen als solche kenntlich gemacht habe.<br />

<strong>Rostock</strong>, den 04.05.2007

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!